targeting antioxidant systems to treat pancreatic cancer · immune checkpoint blockade, inhibition...

91
Targeting Antioxidant Systems to Treat Pancreatic Cancer by Francesco Gabriele Tatangelo Fazzari A thesis submitted in conformity with the requirements for the degree of Master of Science Department of Laboratory Medicine & Pathobiology University of Toronto © Copyright by Francesco Gabriele Tatangelo Fazzari 2018

Upload: others

Post on 06-Jun-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

Targeting Antioxidant Systems to Treat Pancreatic Cancer

by

Francesco Gabriele Tatangelo Fazzari

A thesis submitted in conformity with the requirements for the degree of Master of Science

Department of Laboratory Medicine & Pathobiology University of Toronto

© Copyright by Francesco Gabriele Tatangelo Fazzari 2018

Page 2: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

ii

Targeting Antioxidant Systems to Treat Pancreatic Cancer

Francesco Fazzari

Master of Science

Department of Laboratory Medicine & Pathobiology

University of Toronto

2018

Abstract

Pancreatic cancer has the worst prognosis of all cancers. Pancreatic tumours are characterized by

excessive oxidative stress that may provide a unique vulnerability to target in the development of

new therapeutic strategies. I investigated a novel combination therapy targeting the two main

intracellular antioxidant systems, glutathione and thioredoxin. Buthionine sulfoximine (BSO)

targets glutathione synthesis and auranofin targets thioredoxin recycling. These drugs have been

tested in human clinical trials and are strong candidates for repurposing in cancer therapy. I found

that BSO potentiated auranofin’s cytotoxic effects in vitro. Furthermore, the combination

augmented oxidative stress in cells. Auranofin engaged its target TrxR in vitro at concentrations

causing cell death. I also assessed the biodistribution of auranofin in vivo and observed a

preferential accumulation in the stroma 24 hours after treatment. Overall, I conclude that this

treatment strategy can be effective against pancreatic cancer and deserves attention for future

development.

Page 3: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

iii

Acknowledgments

First and foremost, I would like to thank my supervisor, Dr. David Hedley, for providing an

excellent opportunity to research in his lab. His guidance was instrumental in helping me to

develop a successful project and for supporting me as I take the next steps in my career.

I would also like the thank all the members of the Hedley lab past and present for their assistance

throughout my tenure. They provided an enjoyable work environment and were always willing to

lend a helping hand to teach me different techniques.

Next, I would like to thank my committee members, Drs. Richard Hill and Marianne Koritzisnky.

Their critical assessment of my work helped to shape my project to what it has become.

Lastly, I would like to thank my family and friends for providing support throughout my studies,

and for helping me to enjoy my time outside of the lab.

This work was made possible by support through funding from the Ontario Centres of Excellence.

My stipend was partially supported by the University of Toronto Fellowship Award.

Page 4: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

iv

Table of Contents

ACKNOWLEDGMENTS ........................................................................................................... III

TABLE OF CONTENTS ............................................................................................................ IV

LIST OF TABLES .................................................................................................................... VI

LIST OF FIGURES .................................................................................................................. VII

LIST OF ABBREVIATIONS .................................................................................................... VIII

CHAPTER 1 INTRODUCTION .................................................................................................... 1

1.1 PANCREATIC CANCER ............................................................................................................ 1

1.1.1 Pancreatic Cancer Therapy Development ................................................................... 2

1.1.2 Pancreatic Carcinogenesis .......................................................................................... 5

1.2 OXIDATIVE STRESS ................................................................................................................ 6

1.2.1 Hypoxia ....................................................................................................................... 7

1.2.2 The Impact of High ROS Levels .................................................................................... 7

1.2.3 Adaptations to High ROS Levels ................................................................................ 10

1.2.4 The Impact of Low ROS Levels ................................................................................... 13

1.3 THE ROLE OF ANTIOXIDANTS ................................................................................................ 15

1.3.1 The Glutathione Antioxidant System ........................................................................ 16

1.3.2 The Thioredoxin Antioxidant System ........................................................................ 22

1.4 A NEW THERAPEUTIC STRATEGY FOR PANCREATIC CANCER ........................................................ 28

CHAPTER 2 ........................................................................................................................... 30

2.1 INTRODUCTION .................................................................................................................. 31

Page 5: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

v

2.2 MATERIALS & METHODS ..................................................................................................... 33

2.3 RESULTS ........................................................................................................................... 38

2.4 DISCUSSION ...................................................................................................................... 49

CHAPTER 3 ........................................................................................................................... 53

3.1 OVERALL DISCUSSION ......................................................................................................... 54

3.1.1 BSO potentiates the acute cytotoxicity of auranofin ................................................ 54

3.1.2 Apoptosis contributes to cell death .......................................................................... 55

3.1.3 The drugs engage their expected targets ................................................................. 56

3.1.4 Oxidative stress follows combination therapy .......................................................... 56

3.1.5 IMC is an effective tool to examine auranofin in vivo............................................... 57

3.2 FUTURE DIRECTIONS ........................................................................................................... 58

3.3 CONCLUSIONS ................................................................................................................... 60

REFERENCES ........................................................................................................................ 61

Page 6: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

vi

List of Tables

Table 2.1. IC50 values of auranofin monotherapy and in combination with BSO in primary pancreatic cancer cell lines. ............................................................. 39

Table 2.2. Panel of markers measured by IMC using subcutaneous PDX tumours.

...................................................................................................................... 47

Table 2.3. Estimation of the mean ions/pixel in the stroma and epithelium of

subcutaneous tumours from auranofin treated PDX mice. ........................... 49

Page 7: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

vii

List of Figures

Figure 2.1. Dose-response of antioxidant pathway inhibitors, TrxR isozyme levels, and colony formation potential. .................................................................... 40

Figure 2.2. Representative two parameter plots measuring cell death and apoptosis in OCIP23. ..................................................................................... 42

Figure 2.3. Representative two parameter plots assessing the GSH content of apoptotic cells in OCIP23. .............................................................................. 43

Figure 2.4. Effect of treatment on thiol and ROS levels in 9A and OCIP23 cell lines. ...................................................................................................................... 44

Figure 2.5. Auranofin-TrxR interaction in vitro. .................................................... 46

Figure 2.6. Distribution of auranofin in subcutaneous PDX tumours. ................... 48

Page 8: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

viii

List of Abbreviations

5-FU 5-fluorouracil

ARE antioxidant response element

ASK1 apoptosis signaling kinase 1

Bax bcl-2-like protein 4

BCNU carmustine

BSA bovine serum albumin

BSO buthionine sulfoximine

c-FLIP cellular FLICE-inhibitory protein

CETSA cellular thermal shift assay

DCF 2’,7’-dichlorofluorescein

DMEM Dulbecco’s Modified Eagle Medium

DMSO dimethyl sulfoxide

ECM extracellular matrix

EGFR epidermal growth factor receptor

ERK extracellular signal-regulated kinase

FBS fetal bovine serum

FFPE formalin-fixed, paraffin-embedded

G6PD glucose 6-phosphate dehydrogenase

GCL glutamate-cysteine ligase

GCLC glutamate-cysteine ligase catalytic subunit

GCLM glutamate-cysteine ligase modifier subunit

GGT -glutamyltranspeptidase

Page 9: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

ix

GLUD1 glutamate dehydrogenase

GOT1 aspartate transaminase

GPx glutathione peroxidase

GR glutathione reductase

Grx glutaredoxin

GS glutathione synthetase

GSH glutathione

GSSG oxidized glutathione

GST glutathione S-transferase

HNE 4-hydroxynonenal

IKK- inhibitor of nuclear factor kappa-B kinase subunit beta

IMC imaging mass cytometry

Keap1 kelch-like ECH-associated protein 1

mBBr monobromobimane

MDH1 malate dehydrogenase

ME1 malic enzyme 1

MMP mitochondrial membrane potential

Msr methionine sulfoxide reductase

NADPH nicotinamide adenine dinucleotide phosphate

Nrf2 nuclear factor, erythroid derived 2, like 2

P-gp multidrug resistance protein 1

PanIN pancreatic intraepithelial neoplasia

PBS phosphate-buffered saline

PD-L1 programmed death ligand 1

Page 10: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

x

PDX patient-derived xenograft

PEGPH20 pegvorhyaluronidase alfa

PHGDH phosphoglycerate dehydrogenase

PI propidium iodide

PPP pentose phosphate pathway

Prx peroxiredoxin

PTEN phosphatase and tensin homolog

RNR ribonucleotide reductase

ROS reactive oxygen species

RPMI Roswell Park Memorial Institute

RT room temperature

SCID severe combined immunodeficiency

System XC– cystine/glutamate antiporter system

TRAIL TNF-related apoptosis-inducing ligand

Trx thioredoxin

TrxR thioredoxin reductase

VEGF-A vascular endothelial growth factor A

Page 11: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

1

Chapter 1 Introduction

1.1 Pancreatic Cancer

Pancreatic cancer has the worst prognosis amongst all cancers; 5 year survival rates for the disease

range from less than 5 to 8%1–3. Although it is only the twelfth most commonly diagnosed, it is

the fourth most common cause of cancer death in Canada2. Unfortunately, mortality from

pancreatic cancer has not improved during the era of modern medicine4,5. Estimates predict that

pancreatic cancer will become the second most common cause of cancer death behind non-small-

cell lung cancer by 20306. The dismal outlook for patients with pancreatic cancer can be partially

attributed to the lack of progress made in the development of diagnostics and therapeutics, which

have remained stagnant over the last 50 years7. During this time vast improvements have been

made for the diagnosis and treatment of other tumour types. For example, the development of the

Pap smear can reduce cervical cancer death by up to 80%8, and the development of trastuzumab

has increased overall survival by 33% in HER2+ breast cancers9,10. The identification of new

biomarkers and treatment strategies is required to overcome pancreatic cancer.

Early stage pancreatic cancer is generally asymptomatic, leading to difficulties in detection. The

disease may be identified secondary to non-specific symptoms, or as an incidental finding. In

addition to this, screening tests have not been developed. Currently the only treatment that offers

the potential for long-term survival is surgery in early-stage disease, however the 5 year survival

rate in patients with resectable tumours remains only 15-20%11. For patients with advanced

metastatic disease there are few systemic options that offer little benefit in survival.

Page 12: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

2

1.1.1 Pancreatic Cancer Therapy Development

The first systemic therapy used in the treatment of pancreatic cancer was 5-fluorouracil (5-FU). 5-

FU is a non-specific agent that inhibits the synthesis of the nucleoside thymidine. 5-FU provided

a 4 month survival benefit when used in combination with radiotherapy, compared to radiotherapy

alone12. Since the indroduction of 5-FU in pancreatic cancer clinics some small advancements

have been made. Gemcitabine became the new standard in the late 1990s when a clinical trial

showed it offered a significant improvement in median overall survival when compared to 5-FU

(5.6 months vs. 4.4 months)13. However, gemcitabine is another non-specific agent. It is a

nucleoside analog that replaces cytidine during DNA replication. Since the introduction of

gemcitabine in the treatment of pancreatic cancer, two therapies have been developed for clinical

use. The combination therapy FOLFIRINOX (folinic acid, 5-FU, irinotecan, and oxaliplatin)

improved median overall survival compared to gemcitabine in a clinical trial (11.1 months vs. 6.8

months)14, and the combination of nab-paclitaxel with gemcitabine improved median overall

survival compared to gemcitabine alone (8.5 months vs. 6.7 months)15. Although there appears to

be a survival difference between FOLFIRINOX and gemcitabine/nab-paclitaxel, the

FOLFIRINOX regimen is more toxic and the patient population was skewed slightly in favour of

the FOLFIRINOX cohort, leading many investigators to consider the two regimens to be similar16–

18. Although these therapies are an improvement on past therapies, median survival falls short of

one year with either regimen.

Great interest has centred on the development of targeted therapies to provide effective non-toxic

treatment for pancreatic cancer. There have been many proposed targets and clinical trials19,

however no such therapy has entered clinical practice in Canada to date. Trials have examined

agents such as bevacizumab, inhibiting angiogenesis by targeting vascular endothelial growth

Page 13: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

3

factor A (VEGF-A), or cetuximab, inhibiting epidermal growth factor receptor (EGFR), in

combination with gemcitabine to no benefit20,21, among many other attempts. From the trials, only

erlotinib in combination with gemcitabine has shown a significant effect on overall survival (6.2

months vs. 5.9 months), although this is not clinically relevant22. In summary, current attempts at

targeted therapy have proven ineffective in the management of pancreatic cancer. This may change

in the future, however current prospects seem bleak and new biomarkers may be required to

develop these therapies further23.

Beyond targeted therapy there have been attempts to develop immunotherapies for pancreatic

cancer. Immunotherapy is a treatment approach that relies on augmenting the host’s immune

response against cancer24. This method derives from the observation that the immune system can

eliminate malignant cells during the early stages of transformation25. While there have been many

unequivocal successes in other tumour types, particularly in advanced melanoma26–28, pancreatic

cancer has proven resistant to these approaches. Clinical trials have observed a lack of response to

immune checkpoint blockade with ipilimumab29 and programmed death ligand 1 (PD-L1) directed

monoclonal antibodies30. These failures have resulted from the poorly immunogenic and

immunosuppressive nature of the pancreatic cancer microenvironment7,31. A small number of

patients enrolled in these trials have received some therapeutic benefit, leading to continued

interest in developing new immunotherapy strategies. One such strategy includes attempting to

achieve a synergistic response through the use of multiple agents. This may include combining

immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells,

and/or conventional cytotoxic therapies32.

In relation to immunotherapy, other strategies attempt to target additional features of the tumour

microenvironment. Attention has focused on the role of the stroma in determining therapy

Page 14: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

4

response. These tumours are characterized by a dense desmoplastic reaction comprised of fibrotic

tissue, extracellular matrix proteins, pancreatic stellate cells, immune cells, and blood vessels7.

The extensiveness of desmoplasia correlates with worse disease outcome33. Furthermore,

desmoplasia limits tumour vascularization, creating a barrier to drug uptake by tumours34. A

clinical trial using the modified hyaluronidase molecule pegvorhyaluronidase alfa (PEGPH20) to

degrade stromal hyaluronan in combination with nab-paclitaxel and gemcitabine has shown

improvement in progression free survival by the perceived mechanism of increasing tumour

vasculature to aid drug delivery35. However, there is evidence to suggest that the stroma may have

a role in protection against tumour progression and an increased tumour vasculature may

contribute to accelerated growth if not targeted properly7. Therefore, there is a difficult balance to

achieve via this strategy. Clinical trials have also given attention to hypoxia, which is a feature of

solid tumours that can result from the lack of vasculature in the stroma.

Hypoxia is a tumour microenvironment state where oxygen utilization exceeds its supply.

Traditionally, pancreatic tumours have been considered to be characterized by extensive hypoxia,

which may be caused by the lack of blood supply resulting from the desmoplastic reaction36. In

patient-derived xenograft (PDX) mouse models, hypoxia correlates with aggressive growth and

metastasis37, making it a logical choice to target. The prodrug evofosfamide is activated under

hypoxic conditions to release a DNA alkylating agent that inhibits cancer cell replication38.

However, a phase III clinical trial observed no significant improvement in median overall survival

using evofosfamide/gemcitabine compared to gemcitabine alone (8.7 months vs. 7.6 months)38. It

is possible that hypoxia in patient tumours is more variable than previously believed, which would

limit the effectiveness of therapy across the entire patient population39. This can be addressed by

non-invasive imaging using positron-emitting hypoxia tracers in order to stratify patients towards

hypoxia-targeting clinical trials40.

Page 15: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

5

In summary, the most effective systemic therapies available for metastatic pancreatic cancer

provide only a small benefit to patients. Although there are promising avenues to explore further,

attempts to develop new strategies against pancreatic cancer have fallen short of expectations.

Further studies are required to identify and examine novel, exploitable targets in pancreatic cancer.

1.1.2 Pancreatic Carcinogenesis

Four driver genes are frequently mutated in pancreatic cancer. These are the oncogene KRAS and

the tumour suppressors CDKN2A, SMAD4, and TP5341–44. These mutations are shared by

pancreatic intraepithelial neoplasia (PanIN), which is a precursor lesion to pancreatic cancer.

Sequencing of these genes in PanINs suggests a stepwise progression of pancreatic cancer.

Activation of KRAS is present in the earliest PanIN-1 lesions, making it the initiating event in

pancreatic cancer tumorigenesis. Inactivation of CDKN2A is found in PanIN-2, followed by

inactivation of SMAD4 and TP53 in PanIN-3 lesions45. However, more recent whole-genome

sequencing of tumour cell-enriched primaries showed cases that displayed either simultaneous

inactivation of multiple drivers or tumours that did not follow the predicted mutation sequence.

These data are consistent with punctuated rather than gradual evolution46. Regardless, KRAS

activation occurs in more than 90% of human pancreatic cancer and its effects likely have the

greatest impact on pancreatic cancer41. Interestingly, KRAS has been implicated in mediating

pancreatic carcinogenesis in part through its impact on oxidative stress. Its most important effect

on oxidative stress is the promotion of low intracellular reactive oxygen species (ROS) levels by

inducing nuclear factor, erythroid derived 2, like 2 (Nrf2) activation47. This aspect will be

discussed further in the next section.

Page 16: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

6

1.2 Oxidative Stress

Pancreatic cancer, among other cancers, is characterized by excessive oxidative stress48. Oxidative

stress is caused by the overproduction of ROS, leading to damage of lipids, proteins, and DNA49.

It occurs when there is an imbalance between intracellular ROS and antioxidants in favour of the

oxidants. ROS are produced at an extreme rate due to the high metabolism inherent in transformed

cells, and may promote aggressive tumour features50,51. Oxidative stress and ROS represent an

interesting target in pancreatic cancer because ROS exhibit dual roles; lower concentrations of

ROS can stimulate cancer cell proliferation, whereas higher concentrations of ROS are damaging

to cells49,52. Cancer cells adapt during oxidative stress conditions to maintain low ROS levels to

avoid cell death and promote proliferation by enhancing antioxidant defences53.

ROS exist as multiple species including superoxide (O2•–) , hydroxyl radical (HO•), and hydrogen

peroxide (H2O2)54. They are produced in various subcellular compartments: the cytoplasm, the

mitochondria, peroxisomes, and the endoplasmic reticulum. The major source of cytoplasmic ROS

in pancreatic cancer is NADPH oxidase, which produces superoxide anion by facilitating the

reduction of molecular oxygen by NADPH55. Mitochondrial ROS are produced in dysfunctional

mitochondria that result from inefficient hypermetabolism in rapidly proliferating cancer cells.

This can cause leakage of electrons at complex I and complex III of the mitochondrial electron

transport chain, leading to the production of superoxide and its dismutation to hydrogen peroxide

by superoxide dismutases56,57. Peroxisomal ROS are produced during the -oxidation of fatty

acids58. Lastly, endoplasmic reticulum can contribute to ROS production through its role in

providing an oxidizing environment to promote disulfide bonding and protein folding59. Although

ROS themselves are short-lived molecules, they can generate longer lasting lipid peroxides such

as 4-hydroxynonenal (HNE)60. HNE is capable of reacting with cellular macromolecules such as

Page 17: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

7

proteins, lipids, and DNA. In addition, HNE can stimulate signaling cascades to affect cell

proliferation and differentiation61. As such, it has been implicated in different cancers.

1.2.1 Hypoxia

Hypoxic tumours experience both chronic and acute hypoxia. Chronic hypoxia exists in regions

located away from blood vessels and is limited by the diffusion of oxygen through the dense

tumour stroma. Acute or cycling hypoxia exists in regions proximal to blood vessels and is

perfusion-limited in that there are temporal instabilities in blood flow through a tumour. Cycling

hypoxia generates greater oxidative stress because the acute influx of oxygen can contribute to an

increase in ROS62. In pancreatic xenografts hypoxia is associated with aggressive growth and

metastasis37. However, human pancreatic tumours may not possess extensive hypoxia as

previously suggested40. Although hypoxia stems from a lack of oxygen, hypoxic tumours have

been stated to overproduce ROS54. However, direct evidence suggesting this phenomenon is

lacking. Hypoxia can contribute to resistance to chemo and radiotherapy, which rely on ROS to

mediate their effects63. This evidence contradicts the idea hypoxia of greater oxidative stress

during hypoxia.

1.2.2 The Impact of High ROS Levels

At high levels ROS disrupt cellular processes by non-specifically damaging proteins, lipids, and

DNA. Furthermore, ROS are implicated in both the extrinsic (death receptor-dependent) and

intrinsic (mitochondrial) pathways of apoptosis. The extrinsic pathway is mediated by death-

inducing ligands, such as TNF-related apoptosis-inducing ligand (TRAIL)64. Following ligand-

receptor interaction the caspase cascade is activated, leading to apoptosis65. ROS negatively

regulate cellular FLICE-inhibitory protein (c-FLIP) by a post-transcriptional method to target it

for degradation and activate the extrinsic apoptosis pathway. The intrinsic pathway is mediated by

Page 18: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

8

a permeability transition in the mitochondria. Mitochondrial stress, including oxidative stress,

stimulates pore formation in the inner mitochondrial membrane which causes cytochrome-c

mobilization and mitochondrial swelling66. This can lead to rupture of the outer membrane, and

with the help of pro or anti-apoptotic Bcl-2 family member proteins can stimulate or prevent

apoptosis. Bcl-2-like protein 4 (Bax) in particular mediates the formation of the mitochondrial

apoptosis-induced channel (MAC) to release mitochondrial pro-apoptotic factors, such as

cytochrome-c, which leads to activation of the executioner caspase proteins67–69. As mitochondria

are a significant source of ROS production, the intrinsic pathway is the largest contributor to ROS-

mediated apoptosis. In this pathway, electron transport chain-produced superoxide triggers

eventual cytochrome-c release, leading to the induction of apoptosis70. Apoptosis following

oxidative stress displays several distinct steps that can be measured by flow cytometry methods.

Following glutathione (GSH) depletion in cells, ROS production causes the mitochondrial

permeability transition and the subsequent loss of mitochondrial membrane potential (MMP), due

to cellular damage. Finally, cells lose viability after perforation of the cell membrane71.

In addition to apoptosis, ROS contribute to ferroptosis, a form of cell death that depends on

intracellular iron72. NADPH oxidase-induced accumulation of ROS was identified to contribute to

ferroptosis. Furthermore, the compounds erastin and sulfasalazine, which inhibit the

cystine/glutamate antiporter system (system XC–), lead to toxic accumulation of ROS and

ferroptosis72,73. The lethal effect can be reversed by addition of the lipophilic antioxidants

ferrostatin-1 and vitamin E, suggesting an important role for lipid radicals in mediating

ferroptosis72,74. KRAS driven cancer cells have been shown to be susceptible to ROS-mediated

ferroptosis following treatment with lanperisone, an FDA-approved muscle relaxant75, suggesting

that ROS can be utilized to induce ferroptosis in a KRAS driven tumour such as pancreatic cancer.

Page 19: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

9

High ROS also have indirect effects on cell death by sensitizing cancer cells to chemo and

radiotherapy. The multidrug resistance protein 1 (P-gp) facilitates the removal of several

anticancer agents from cells. High ROS concentrations have been associated with decreased P-gp

expression, which allows for cytotoxic drugs to accumulate within a cell76. Furthermore, ROS have

been identified to mediate the effects of some cytotoxic chemotherapies. For example, gemcitabine

acts through a ROS-dependent mechanism that may promote metastasis due to the action of

antioxidants77. This feature may contribute to gemcitabine’s limited effectiveness in pancreatic

cancer. Moreover, both 5-FU and oxaliplatin have been observed to augment oxidative stress in

patients with colon cancer78. These drugs are two of the components of the combination

FOLFIRINOX, which is one of the most effective systemic therapies against metastatic pancreatic

cancer. Paclitaxel, which is the main component of the nab-paclitaxel/gemcitabine therapy for

pancreatic cancer, and doxorubicin also enhance oxidative stress79,80. While augmented oxidative

stress is not the primary mechanism of action of these therapies, it is interesting to note that the

most effective therapies in clinical use for pancreatic cancer elicit an effect on the redox

environment of the tumour.

In addition to chemosensitization, high ROS levels are involved in mediating the cytotoxic effects

of radiotherapy. Radiotherapy utilizes ionizing radiation to generate ROS, particularly hydroxyl

radical. Hydroxyl radicals enact the therapeutic and side effects of radiotherapy. It propagates by

reacting with other molecules including DNA, leading to abundant DNA lesions and subsequent

cell death81. It has been observed that radical scavengers are effective at attenuating the cytotoxic

effects of radiation, indicating that ROS are required82–84. The evidence presenting the role of ROS

in chemo and radiosensitization suggests that exploiting an increase in ROS can be effective in the

development of new therapies.

Page 20: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

10

Cancer cells require multiple adaptations to protect from inherently high ROS levels allowing them

to survive and proliferate. Signaling through the oxidative stress response mediates this adaptation.

The master regulator of this response is the transcription factor Nrf2. Nrf2’s activity is modulated

by its repressor Kelch-like ECH-associated protein 1 (Keap1). In a normal redox state Nrf2 is

constitutively degraded. Keap1 promotes its ubiquitination by Cullin E3 ubiquitin ligase to label

Nrf2 for proteasome degradation85. During oxidative stress conditions, redox-sensitive cysteine

residues on Keap1 are oxidized by ROS. In this state Keap1 is unable to bind Nrf2 to recruit

ubiquitin ligases to target it for degradation. Nrf2 subsequently translocates to the nucleus where

it heterodimerizes with small Maf proteins to stimulate transcription of genes associated with

antioxidant response element (ARE) promoters86. The lipid peroxide HNE is a potent inducer of

Nrf2 by covalent binding to Keap1 redox-sensitive cysteines and by activating upstream kinases

such as protein kinase C87,88 It has been observed that nuclear Nrf2 expression associates with poor

survival in pancreatic cancer89, whereas cytoplasmic Keap1 expression associates with reduced

metastasis and membrane-associated Keap1 displays a protective effect on survival90.

1.2.3 Adaptations to High ROS Levels

Nrf2 and ARE positively regulate the transcription of several cytoprotective genes including drug-

metabolizing enzymes, drug transporters, and antioxidants91. Nrf2 regulates the expression of

many phase I and II drug-metabolizing enzymes92–94, as well as members of the multi-drug

resistance-associated gene family of multidrug transporters95. It can indirectly modulate ROS

levels by regulating free Fe (II) homeostasis. This prevents the Fenton reaction, where Fe(II)

catalyses the oxidation of hydrogen peroxide to the more reactive hydroxyl radical96. This function

occurs by promoting the release of Fe (II) from heme, and its subsequent sequestration. However,

Nrf2’s most important role is functioning as the master regulator of antioxidants. The production

Page 21: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

11

and regeneration of the most abundant antioxidant cofactor, GSH, is controlled exclusively by

Nrf297. In addition, Nrf2 supports the expression of enzymes involved in GSH utilization98, as well

as enzymes involved in the production, regeneration, and utilization of thioredoxin (Trx)99. These

actions of Nrf2 are responsible for direct modulation of oxidative stress. These antioxidants will

be discussed in further detail in section 1.3. In parallel to antioxidant production, cancer cells

require the redirection of cellular metabolism to support antioxidant function.

Cancer cell metabolism differs from that of normal cells and requires abundant ATP production to

support rapid proliferation. Cancer cells experience a metabolic switch to rely on aerobic

glycolysis, rather than oxidative phosphorylation, to supplement their needs. This phenomenon is

termed the Warburg effect100,101. Constitutively active KRAS in pancreatic cancer plays a key role

in activating the metabolic switch towards supporting anabolic pathways102. In addition to this

switch, Nrf2 reprograms metabolism in oxidative stress conditions to support the antioxidants it

promotes. It reprograms metabolism by increasing transcription of NADPH-generating enzymes

to modulate glucose, glutamine, and serine metabolism103,104. NADPH is an important cofactor

that is required by antioxidant enzymes to supply reducing potential. Nrf2 promotes a proliferative

phenotype through the PI3K-Akt axis, which redirects glucose into anabolic pathways to increase

metabolism103. Three major pathways that produce NADPH are regulated by Nrf2: the pentose

phosphate pathway (PPP), the serine synthesis and folate pathway, and the malic enzyme pathway

of glutamine metabolism.

The PPP is a metabolic pathway that runs in parallel to glycolysis. It consists of oxidative and non-

oxidative phases. The oxidative arm is responsible for NADPH production, while the non-

oxidative arm is responsible for the generation of ribonucleotides105. Both phases are critical to

cancer cell survival. Glucose 6-phosphate dehydrogenase (G6PD) is the first enzyme in the

Page 22: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

12

oxidative arm of the PPP and is positively regulated by Nrf293. G6PD commits glucose 6-

phosphate to the PPP, rather than glycolysis105. The PPP has been observed to promote cell survival

and proliferation during neoplastic transformation106. Although the oxidative phase has generally

been implicated to supply cancers with NADPH, constitutively active KRAS has been observed

to uncouple ribose biogenesis from NADPH production in pancreatic cancer107. Glycolytic

intermediates are utilized specifically in the non-oxidative arm of the PPP, but not the oxidative

arm. Inhibition of the PPP by knockdown of G6PD, as well as glucose deprivation, in pancreatic

cancer cells led to minimal effect on ROS, indicating that the oxidative arm is not required to

maintain ROS levels in pancreatic cancer108. ROS could be affected, however, by modulating the

other two NADPH production pathways.

The serine synthesis pathway is another glycolytic shunt responsible for NADPH production. In

this pathway the glycolytic intermediate 3-phosphoglycerate is converted to serine via three

reactions. The first committed step is catalyzed by phosphoglycerate dehydrogenase (PHGDH).

After its production serine can be used in the folate cycle to generate glycine and NADPH109. The

transcription of the enzymes involved in this process is regulated by Nrf2104. Overexpression of

PHGDH has been reported in non-small cell lung, colorectal, breast, and pancreatic cancers104,110–

112. In pancreatic cancer, PHGDH has been described as an independent prognostic indicator for

patients; increased expression is associated with poor prognosis. Furthermore, PHGDH promotes

proliferation, migration, and invasion of pancreatic cancer cells in vitro112. Reduced levels of

serine hydroxymethyltransferase, the enzyme responsible for preparing serine for the folate cycle,

were responsible for decreasing NADPH while increasing ROS113. Therefore, Nrf2 promotes the

serine synthesis and folate pathway to produce NADPH, which allows the cancer cells to adapt to

oxidative stress and may contribute to worse outcomes in pancreatic cancer.

Page 23: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

13

Finally, glutamine metabolism through the malic enzyme (ME1) pathway is another source of

NADPH production that may be the most important in pancreatic cancer108. ME1 is positively

regulated by Nrf299. Cancer cells exhibit glutamine addiction because of its essential canonical

roles in supplying intermediates for the Kreb’s cycle through anaplerosis, and its roles in

nucleotide, amino acid, and lipid biosynthesis114–116. Most cells utilize the enzyme glutamate

dehydrogenase (GLUD1) to convert glutamine-derived glutamate into -ketoglutarate to fuel the

Krebs cycle, however constitutively active KRAS in pancreatic cancer downregulates GLUD1 and

upregulates aspartate transaminase (GOT1) to shunt glutamine into the non-canonical ME1

pathway for NADPH production108. In this pathway glutamine-derived aspartate is converted to

oxaloacetate via GOT1 in the cytoplasm, and then oxaloacetate is metabolized by malate

dehydrogenase (MDH1) into malate. The final step involves the oxidation of malate to pyruvate

by ME1. This reaction is responsible for NADPH regeneration and assists in protecting from

damage during oxidative stress117. Confirming this, glutamine deprivation and ME1 knockdown

led to elevated ROS and decreased clonogenicity in pancreatic cancer cells. The cells could be

rescued with supplemented oxaloacetate, malate, or GSH, indicating that glutamine is required for

redox homeostasis in pancreatic cancer108. Further demonstrating the importance of glutamine

metabolism in pancreatic cancer, ME1 pathway inhibition diminished tumour xenograft growth108.

1.2.4 The Impact of Low ROS Levels

The adaptations controlled by Nrf2 maintain ROS at low to moderate levels (ie. above background

levels but below cytotoxic levels) to prevent cell death. These low ROS levels are implicated in

stem cell phenotype and in tumour cell proliferation and act as important signaling molecules

capable of regulating various functions including proliferation, apoptosis, and tumour cell

invasion48. Low ROS levels are associated with maintaining cancer stem cell-like properties in

Page 24: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

14

cells that confers therapy resistance118. For example, human breast cancer stem cells possess robust

ROS scavenging systems that maintain low ROS levels84. This feature prevents ROS-induced

DNA damage following ionizing radiation, protecting the cells from death. Furthermore,

pharmacologic depletion of antioxidants results in loss of stem cell properties84.

Low ROS are involved in signal transduction pathways to regulate various cellular responses. This

function is made possible because of ROS-mediated inactivation of protein-tyrosine phosphatases,

which allows for increased signaling through receptor tyrosine kinases119. As such, ROS activate

signaling of the MAPK pathway by activating p38 MAPK to promote tumour cell differentiation

and proliferation120. In addition, ROS can inactivate PI3K/Akt phosphatases like PTEN. This

stimulates signaling of the PI3K pathway to promote proliferation and apoptosis evasion55,121.

Proliferation promotion by ROS has been observed in various cancer cell types122,123. These

functions implicate ROS in carcinogenesis. Moreover, ROS may contribute to the promotion of

invasion and metastasis. Tumour cell invasion and metastasis requires cell mobility through the

extracellular matrix (ECM). Cells degrade glycosaminoglycan in the ECM to allow for motility.

ROS-dependent degradation of the ECM occurs due to the stimulation of matrix metalloproteinase

activity, expression, and secretion124. ROS-activated MAPK pathways contribute to the expression

of matrix metalloproteinases. Specifically, extracellular signal-regulated kinase (ERK) activation

by ROS promotes invasion and metastasis125. The metastatic phenotype is enabled and sustained

by concomitant ROS production and the ability to protect from the cytotoxicity of oxidative stress

with antioxidants to maintain low ROS levels126. These functions of low ROS levels allow for

tumour growth and spread that may contribute to worse prognosis in pancreatic cancer.

Taken together this information indicates that low ROS levels contribute to many of the

tumorigenic properties of cancer cells. Cancer cells utilize antioxidants to protect themselves from

Page 25: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

15

ROS and maintain these low ROS levels. These antioxidants are regulated by the master regulator

Nrf2 during the oxidative stress response. ROS scavenging prevents ROS-induced toxicity and

assists in ROS-mediated proliferation. Strategies designed to exacerbate ROS through targeted

inhibition of antioxidants may provide a relatively non-toxic therapy, as cancer cells are reliant on

antioxidants whereas normal cells can survive change. This strategy has been employed using

various drugs against a number of targets to great success75,127,128.

1.3 The Role of Antioxidants

The most important factor in the cell that moderates ROS levels are the antioxidants regulated by

Nrf2129. The role of antioxidants and Nrf2 in cancer is controversial as studies suggest them to be

either tumour suppressive or oncogenic. The answer to this controversy is likely to be that the role

of antioxidants is context specific91,130. Nrf2-knockout mice are more susceptible to carcinogenesis

and show diminished efficacy of chemopreventive enzymes131. Furthermore, the natural

compound sulforaphane activates Nrf2 and was found in a phase II clinical trial to protect against

hepatocellular carcinoma by preventing DNA adduct formation132. Although these studies suggest

that the antioxidant response mediated by Nrf2 is required for chemoprevention, most clinical trials

suggest that dietary antioxidants do not provide this effect133,134, and may even increase cancer

risk135,136.

Nrf2 may promote tumorigenesis by protecting cells from oxidative stress and maintaining low

ROS. Oncogenes such as KRAS in pancreatic cancer increase Nrf2 to provide cytoprotection and

decrease ROS levels47. Loss-of-function mutations in KEAP1, resulting in constitutive Nrf2

activation, have been identified in various solid tumours137. Lastly, high Nrf2 expression has been

correlated with resistance to platinum-based chemotherapy in ovarian cancer138.

Page 26: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

16

These contrasting sets of data suggest that the role of Nrf2 may depend on the stage of

tumorigenesis. Before the cancer development, Nrf2 and antioxidants may help to suppress cancer

by the prevention of mutations and oxidative stress. Once the cancer develops, Nrf2 and the

antioxidants it regulates may promote cancer progression91,130. Therefore, once a tumour has been

developed, antioxidants may prove to be an important target for cancer therapy. As previously

discussed, the GSH and Trx systems are both coordinated by Nrf2. These are likely two of the

main factors contributing to tumour cytoprotection. Not only is the production of GSH and Trx

controlled by Nrf2, but also the production of the enzymes that recycle and utilize them to evoke

their effects.

1.3.1 The Glutathione Antioxidant System

GSH is the most abundant non-protein thiol in cells and is critical to providing protection from

oxidative stress139. GSH exists in two forms: reduced GSH, or oxidized glutathione (GSSG) which

is a dimer. A number of enzymes responsible for GSH homeostasis are regulated by Nrf2 during

periods of oxidative stress. GSH synthesis is performed in a two-step process from the amino acids

cysteine, glutamate, and glycine. The first reaction in GSH biosynthesis is the rate-limiting step

catalyzed by glutamate-cysteine ligase (GCL), which has both a catalytic (GCLC) and modifier

(GCLM) subunit140. GCLC performs the catalytic activity of the enzyme141, while GCLM is an

important regulatory subunit that lowers the Km for glutamate142. In this reaction glutamate and

cysteine are ligated to form -glutamylcysteine. This dimer contains a non-conventional peptide

bond between the -carboxyl group of glutamate and the amine of cysteine, rather than the -

carboxyl group. Hydrolysis of this bond can only be performed by -glutamyltranspeptidase

(GGT) which is found on the external surface of certain cell types and supplies cells with GSH

precursors143. As such GSH resists intracellular degradation. The second reaction to add glycine

Page 27: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

17

to -glutamylcysteine is catalyzed by GSH synthetase (GS). Overexpression of GS fails to increase

GSH levels, whereas GCL overexpression leads to an increase in GSH144, supporting the notion

that GCL is more important in the regulation of GSH synthesis. Since -glutamylcysteine is found

at very low intracellular concentrations, GCL is considered to be rate-limiting140. Cystine uptake

through the system XC– transporter can also be a limiting step in GSH synthesis, as intracellular

cysteine concentrations are low145. Cysteine auto-oxidizes extracellularly to cystine, which can

then be taken up by the system XC– transporter before rapid intracellular reduction145. The

expression of both GCL subunits and the critical xCT subunit of the system XC– are regulated by

Nrf2146,147.

1.3.1.1 Functions of the Glutathione System

As a redox sensitive thiol the primary function of GSH is to scavenge ROS148. Supporting the

antioxidant role of GSH, depletion of GSH leads to increases in ROS and subsequent ROS-induced

toxicity149. GSH is particularly required in the mitochondria to protect from pathologically

generated oxidative stress150. This antioxidant function is performed in reduction reactions

catalyzed by glutathione peroxidase (GPx). GPx catalyzes the reduction of hydrogen peroxide and

lipid peroxides, which in turn oxidizes GSH to GSSG151. Members of the GPx enzyme family,

including GPx2 and GPx4, are regulated by Nrf294,152. GSH recycling is performed when cells

reduce GSSG using the enzyme glutathione reductase (GR) and NADPH to maintain a reducing

environment. This key enzyme for maintaining GSH homeostasis is regulated by Nrf2153. Since

the GSH to GSSG ratio is a significant contributor to the redox equilibrium, cells require

substantially more GSH than GSSG to maintain a reducing environment. During periods of

oxidative stress when GSSG accumulates, cells can export GSSG when the capacity of GR to

recycle GSH is outmatched154. Another mechanism to prevent toxic GSSG accumulation during

Page 28: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

18

oxidative stress is the reaction of GSSG with a protein thiol catalyzed by protein disulfide

isomerase to form a protein mixed disulfide155.

GSH can contribute to redox signaling in addition to its antioxidant function. This role is executed

by modifying the oxidation state of critical modifying cysteine residues in proteins156.

Mechanistically this is performed by the reversible binding of GSH to protein thiols to alter the

protein structure and function. This process is called glutathionylation and can activate or

inactivate proteins157. Glutathionylation can happen enzymatically with the assistance of GSH S-

transferase (GST) or non-enzymatically158,159. This mechanism exists to protect sensitive proteins

from irreversible oxidation. Deglutathionylation is catalyzed by the oxidoreductase glutaredoxin

(Grx) which is itself reduced by GSH160. A significant number of proteins critical in cell signaling

have been identified to be regulated by glutathionylation. These include GAPDH, inhibitor of

nuclear factor kappa-B kinase subunit beta (IKK-), and p53161–164. This process may affect ROS

generation, as glutathionylation of electron transport chain proteins can alter their function165. GST

has an additional role in phase II detoxification of xenobiotics, including chemotherapeutic drugs.

In this process GSH is conjugated to xenobiotics during metabolism to prepare for excretion166.

GST-mediated detoxification can contribute to chemoresistance167.

GSH contributes to the regulation of growth in both normal and transformed cells. Increased GSH

levels are essential for cell cycle progression and are associated with proliferation168–170.

Hepatocytes increase GSH in response to injury, as has been observed during liver regeneration

following partial hepatectomy171. Blocking this GSH generation causes a reduction in the total

amount of DNA synthesis associated with cell division170. GSH levels in various cancer types

including hepatocellular carcinoma, melanoma, and pancreatic cancer have been correlated with

cell and tumour growth170,172,173. Interestingly, the redox environment of the nucleus has been

Page 29: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

19

observed to be critical during the cell cycle. During proliferation GSH localizes to the nucleus and

can affect gene expression of multiple proteins through influence on histones174. Cell cycle

progression requires a reducing environment moderated by GSH to occur. In addition to cell

growth regulation, GSH is also involved in cell death regulation. This regulation is carried out

through effects on the mechanism of apoptosis. GSH levels impact the expression and activity of

caspases that are involved in effecting apoptosis175. As previously mentioned, GSH levels fall as

ROS propagate during the apoptotic cascade149. This can occur because of decreased GCL activity

following GCLC cleavage by caspase 3 during apoptosis176.

1.3.1.2 Targeting the Glutathione System in Cancer

GSH is critical for malignant behaviour of cancer cells. GSH is required during tumour initiation,

as its depletion impairs tumorigenesis127. This is because GSH prevents oxidative stress from

enacting its cytotoxic effects on transformed cells. Following tumour formation GSH becomes

important to maintain high growth rates, as GSH levels are correlated to growth177. These levels

are elevated in cancer due to increased expression of GCL and the system XC–178,179. In addition,

GGT is an early marker of neoplastic transformation and assists with the GSH cycle to supply

cancer cells with GSH metabolic precursors180. GGT overexpression has been observed in

melanoma and hepatocellular carcinoma181,182, among other cancers. Lastly, GSH is required for

establishing distant metastases because metastatic cells experience higher oxidative stress. In

response they elevate GSH levels to combat the microenvironment and maintain redox

homeostasis126. To highlight the importance of GSH in cancer, many studies have attempted

pharmacologic GSH depletion and observed sensitization to various therapies. Cancer cells are

sensitized to radiation upon GSH depletion84. In addition, chemoresistance has been associated

with high GSH levels in multiple myeloma and ovarian cancer165,166. This chemoresistance can be

eliminated following depletion of GSH184

Page 30: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

20

As GSH is critical to cancer progression and survival there is reason to believe that it can be

exploited in cancer therapy. Various key nodes in GSH metabolism can be targeted as an anticancer

strategy. First, substrate supply for GSH biosynthesis can be targeted. Glutamine analogs targeting

GGT have been evaluated in humans but have been found too toxic for clinical use185. Newer

uncompetitive inhibitors are in development but have yet to prove effective and will require time

before clinical application186. The system XC– transporter may be a critical target of an anti-GSH

cancer therapy as it is required to regulate intracellular cysteine and has been observed to be

necessary for cancer cell survival187. Multiple inhibitors against this system have been described.

Sulfasalazine is a drug that has been used in clinics for rheumatoid arthritis, ulcerative colitis, and

Crohn’s disease, and inhibits the system XC–188. However, the mechanism of action is poorly

understood, it is metabolically unstable, and lacks potency73. Erastin is a newer molecule that is

more potent than sulfasalazine and more selective for system XC–74. However, it has yet to be tested

in humans.

A second potential target critical to GSH metabolism is GSH recycling. The nitrogen mustard

compound carmustine (BCNU) is primarily an alkylating agent used in chemotherapy for brain

cancer and lymphomas189,190, however its off-target effect is GR inhibition191. Unfortunately,

BCNU is in short supply, costly, and displays prolonged myelosuppression and potentially lethal

pulmonary toxicity190,192,193. A newer molecule 2-AAPA has been described as a GR inhibitor that

increases GSSG and stimulates protein glutathionylation194. However, 2-AAPA has yet to be tested

in humans and possesses off-target effects including Grx and thioredoxin reductase (TrxR)

inhibition195,196. These off-target effects and potentially others that have yet to be uncovered make

2-AAPA too unpredictable for use in cancer therapy.

Page 31: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

21

Lastly, a third strategy involves direct targeting of GSH synthesis. GCL, the rate-limiting enzyme,

is more important for GSH levels than GS, the enzyme catalyzing the second step of GSH

synthesis144. Therefore, GCL would be the ideal target. Fortunately, an irreversible inhibitor of

GCL, buthionine sulfoximine (BSO), has been studied extensively in vitro, in vivo, and in human

clinical trials127,197,198. BSO has proven to be minimally toxic in humans when used on its own199.

Although it is rapidly metabolized following intravenous bolus, concentrations of ~500M

following continuous infusion are achievable in humans while maintaining safe toxicity profiles200.

BSO has garnered great interest for use in drug combinations against cancers that have been

observed to require GSH for chemoresistance197. Mechanistically, it contributes to oxidative stress

in the cells by acute GSH depletion and long-term ROS exacerbation149. This can lead to both

apoptosis and ferroptosis149,201. BSO’s safety profile, specificity, low cost, previous clinical trials,

and mechanistic understanding make it an excellent drug to develop further for cancer therapy that

modulates oxidative stress.

BSO on its own is not very cytotoxic with short-term treatments and many cell lines are resistant

to acute GSH depletion197. It is likely that therapy involving BSO would require a second agent to

enhance toxicity. Adaptation to GSH depletion involves Nrf2 activation to upregulate other

antioxidants, including Trx202. Since antioxidants have similar, sometimes overlapping functions,

compensation from alternative antioxidant systems is a plausible mechanism for resistance to GSH

depletion. Compensation between antioxidants has been observed in other contexts including

between GSH and Trx, and between Trx and heme oxygenase127,203,204. Therefore, therapy tailored

to inhibit more than one antioxidant system may be able to overcome compensation and produce

strong anti-cancer effects.

Page 32: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

22

1.3.2 The Thioredoxin Antioxidant System

The Trx system is a reductase system that is required for DNA synthesis and protection from

oxidative stress203. The three major components of the system are Trx, TrxR, and NADPH. As

with other antioxidant genes, Trx and TrxR are transcriptionally regulated during oxidative stress

by Nrf299,205. Trx is a small (~12 kDa) reductase that is the effector of the Trx system. It catalyzes

the reduction of protein disulfides. The active site dithiol in reduced Trx interacts with various

protein substrates via a thiol-disulfide exchange reaction that forms a disulfide in Trx (oxidized

Trx). This oxidized form is reduced by electron transfer from NADPH via the reaction catalyzed

by TrxR. There is a mitochondrial Trx2 with two active site cysteines, and cytosolic and nuclear

Trx1 with 3 additional cysteines that contribute to enzymatic regulation206. TrxR is a homodimeric

oxidoreductase that recycles Trx, and belongs to the same protein family as GR207. Mammalian

cells contain three TrxRs: cytosolic TrxR1, mitochondrial TrxR2, and testis-specific TrxR3208. Of

these, TrxR1 has been identified as a fitness gene209. Although structurally and functionally

similar, the TrxR enzymes have different amino acid sequences210. These enzymes contain a

unique C-terminal selenocysteine residue that contributes to its reaction mechanism.

In the mechanism of TrxR reduction electrons from NADPH are first transferred to enzyme-bound

FAD, then FAD transfers reducing equivalents to the N-terminal disulfide in the redox centre of

the same subunit. This reduces the disulfide to a dithiol, which then transfers the electrons to the

C-terminal selenenylsulfide of the other subunit. Addition of another NADPH molecule reduces

these residues to a selenol-thiol pair211. The reduced enzyme can transfer these electrons to its

substrates, including proteins such as oxidized Trx and Grx2, and small molecules such as

hydrogen peroxide and cytochrome c212. When transferring electrons to Trx the selenol of TrxR is

deprotonated to a selenolate anion. This anion attacks the disulfide of oxidized Trx, resulting in a

Page 33: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

23

TrxR-Trx mixed selenenylsulfide. The free thiol on the C-terminal TrxR reduces this bond, causing

the reformation of the selenenylsulfide in TrxR. This allows the now reduced Trx to leave with

reduced dithiols in its active site211. This mechanism is made possible because of the

selenocysteine residue. Selenocysteine is a cysteine analog that replaces sulfur with selenium. This

maintains many chemical properties of the amino acid, however the pKa of the selenol group in

selenocysteine is significantly lower than the thiol group in cysteine (cysteine pKa = ~5.8 vs.

selenocysteine pKa = ~8.3). This property allows the selenol to exist primarily as the ionized

selenolate at physiological conditions, whereas the thiol exists in the protonated form213. The

ionized molecule is more reactive than the protonated version. In fact, replacement of

selenocysteine with cysteine dramatically decreases activity of TrxR1214.

1.3.2.1 Functions of the Thioredoxin System

As one of the major thiol-dependent electron donor systems in the cell, the Trx system carries out

multiple important functions. The Trx system can perform mechanisms of both direct and indirect

ROS scavenging. First, Trx can directly neutralize ROS. This action involves quenching hydroxyl

radical by providing electrons, and requires Trx’s catalytic cysteines, as shown by loss of ROS

scavenging following mutation of these cysteines to alanine215. Second, TrxR can indirectly

scavenge ROS by providing electrons to other endogenous antioxidants, including ubiquinone.

The reduced form of ubiquinone, ubiquinol, can act as a ROS scavenger by reducing radicals to

protect from lipid peroxidation216. The reduction of ubiquinone is catalyzed by TrxR under

physiologic conditions, and may contribute to protection from oxidative stress217.

A critical function of the Trx system involves the regulation of protein disulfide formation. During

oxidative stress ROS enact damage to cellular macromolecules. As proteins are the most highly

concentrated macromolecule in cells, they are likely the most significant target of ROS218. Not all

Page 34: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

24

proteins are equally susceptible to oxidative damage, as subcellular localization and content of

oxidation-sensitive amino acids may contribute219. The most oxidation-sensitive amino acid

residue is cysteine, that readily forms disulfide bridges with other cysteine residues upon

oxidation220. These aberrant disulfides can cause structural changes that impact protein function.

Trx maintains protein redox status via its reductase activity which can reduce disulfides in an

extensive number of target proteins to dithiols221. Trx targets regulate a vast array of cellular

processes including proliferation, metabolism, and apoptosis. In general, the role of Trx is to

provide a reducing environment to protect from oxidative stress to prevent cell death and promote

cell growth. The most important targets of Trx are reductive enzymes including Prx, ribonucleotide

reductase, and methionine sulfoxide reductase. This highlights the importance of Trx in the

maintenance of other reductive pathways222.

Prxs are a peroxidase family that act primarily as cellular antioxidants. Three subgroups of Prx

exist: 2-cysteine Prx isoforms (Prx1-4), atypical 2-cysteine Prx isoform (Prx5), and 1-cysteine Prx

isoform (Prx6)223. Reduced Trx is the electron donor for Prx1-5 which can then remove ROS such

as hydrogen peroxide, lipid hydroperoxides, and peroxynitrite. Hyperoxidation of Prx occurs from

ROS overexposure. The catalytic cysteine residues oxidize to form disulfide bonds, which must

be reversed by Trx224. This establishes a redox cascade where the flow of electrons travels from

NADPH through the Trx system to Prx and finally to neutralize toxic ROS225. The speed of

hydrogen peroxide scavenging by Prx is comparable to other significant ROS scavenging enzymes

like glutathione peroxidase, on the magnitude of 107 – 108 M-1s-1, indicating that it is a major

contributor to peroxide scavenging in cells226,227. Prxs can inhibit cancer development or promote

tumorigenesis depending on the specific Prx member and cancer context228. However, high levels

of Prxs are often associated with radioresistance and chemoresistance. For example, high Prx2

levels correlates to radioresistance in breast cancer cells, as well as with cisplatin resistance in

Page 35: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

25

gastric cancer cells229,230. Silencing of Prx2 sensitizes colorectal cancer cells to both radiation and

oxaliplatin treatment231. These data indicate that Trx may have an indirect role in promoting

therapy resistance through Prx.

Another significant reductase target of Trx is ribonucleotide reductase (RNR). RNR reduces all

four ribonucleotides to deoxyribonucleotides for use in DNA synthesis. During ribonucleotide

reduction the two active site cysteines of RNR form a double bond which must be reduced by a C-

terminal dithiol through a thiol-disulfide exchange reaction232. The original conformation of the

C-terminal cysteines is restored by an external reductase system, either Trx or Grx which both

have similar catalytic efficiency with mammalian RNR233–235. This redundancy allows for cells to

withstand the loss of one reductase system. This is supported by studies showing that TrxR

downregulation causes no change in deoxyribonucleotide triphosphate pools in mouse cancer

cells236. Some differences may exist as Trx is believed to support S phase electron donation,

whereas Grx is believed to support DNA repair and mitochondrial DNA synthesis235. Interestingly,

Grx is an important effector of the GSH system, supporting compensatory mechanisms between

Trx and GSH.

Methionine is another oxidation-sensitive amino acid. However, protein reduction by Trx is

limited to disulfides. Methionine sulfoxide reductase (Msr) is capable of reducing oxidized

methionine in proteins and acquires its reducing potential from Trx237. Although there are multiple

Msr isoforms with different numbers of catalytic cysteines and mechanisms, they all require Trx

to maintain their reductive capacity222. This is yet another pathway by which Trx is required to

protect cells from damage caused by oxidative stress.

Beyond reductase systems, Trx is responsible for post-translational modifications to regulate

numerous signaling cascades and transcription factors. Apoptosis signaling kinase 1 (ASK1) is

Page 36: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

26

part of the MAPKKK family that activates the JNK and p38 MAP kinase pathway238. Reduced

Trx can bind ASK1 to inhibit its activation, directing it to ubiquitin-mediated degradation, which

prevents ASK1 from activating proapoptotic gene expression239. Inactivation of PTEN, a negative

regulator of the PI3K-Akt pathway, results in the stimulation of proliferation. Trx1 binds PTEN

and inhibits its activity, which may stimulate cancer cell proliferation240. These data indicate a

potential role for Trx in cancer proliferation and highlight the importance of Trx inhibition in

cancer therapy.

1.3.2.2 Targeting the Thioredoxin System in Cancer

A novel strategy for pancreatic cancer therapy may involve Trx system inhibition, as the Trx

system is involved in protection from oxidative stress that may promote cancer. Excluding

downstream targets of Trx, three major targets can be identified within the Trx system: NADPH,

Trx, and TrxR. Since NADPH is ubiquitous and required in a vast array of major biosynthetic

pathways, complete inhibition of NADPH production would not be feasible. The individual

pathways of NADPH production could potentially be targeted. The molecule dichloroacetate

inhibits the oxidative phase of the PPP and has been clinically tested241,242. However, this likely

may not be effective in a KRAS driven tumour like pancreatic cancer since the oxidative phase of

the PPP is less involved in NADPH production than other pathways108. When targeting the serine

biosynthesis pathway PHGDH knockdown in overexpressing cancers inhibits proliferation111.

However, inhibitors of PHGDH, including PKUMDL-WQ-2101 and PKUMDL-WQ-2201, have

only recently been identified and require further development before use in humans243. Targeting

glutamine metabolism through inhibition of ME1 has shown to be promising by inducing

senescence and apoptosis in vitro244. However, an inhibitor of ME1 has yet to be characterized.

An inhibitor of the mitochondrial isoform ME2, NPD389, has been described but not developed245.

Page 37: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

27

As the effector of most functions of the Trx system, Trx would seem to be an ideal target. The

synthetic molecule PX-12 is the most developed Trx inhibitor that covalently binds to Trx cysteine

residues, forming mixed disulfides. PX-12 inhibits breast cancer cell proliferation in vitro and

reduced tumour volume in vivo. Interestingly, PX-12 can be a substrate for TrxR causing

competitive inhibition of TrxR for its normal substrate, oxidized Trx246. PX-12 has been tested in

clinical trials, and was shown to decrease plasma Trx in a phase I trial247. However, a phase II trial

of PX-12 in patients with advanced pancreatic cancer refractory to gemcitabine therapy showed

no antitumour activity248. These results may stem from the inability to achieve adequate plasma

concentrations of PX-12 without significant toxicities because of rapid binding to plasma

components. Combined results of clinical trials suggest that development of PX-12 for intravenous

infusion is not feasible249.

The third and last potential target, TrxR, may contain the most potential for anticancer therapy.

High TrxR1 expression correlates to poor prognosis in hepatocellular carcinoma, breast, and

pancreatic cancer, among others250,251. TrxR1 inhibition following siRNA knockdown

dramatically reduced tumour progression in xenograft models of lung cancer252. In contrast,

normal adult tissues can survive TrxR1 loss253. Recent development has centered around attempts

to develop specific inhibitors of TrxR1 with some success254. However, these molecules require

extensive development and testing before potential clinical use. The pan-TrxR inhibitor auranofin

is an FDA approved drug for the treatment of rheumatoid arthritis that has recently garnered

interest for repurposing in cancer therapy255. Auranofin is an orally available gold complex that

irreversibly inhibits all TrxR isoforms, and is inexpensive, well-characterized, and has relatively

minor toxicities256. Although it has been reported to inhibit proteasome deubiquitinases in addition

to TrxR257, evidence to support this is sparse and seems to be secondary to cell death. Auranofin

is capable of inducing apoptotic cell death following the proliferation of ROS258,259. However, it

Page 38: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

28

appears to work best in combination with other drugs after being testing as a combination therapy

against various cancers260–262. In addition, auranofin has been used in phase I and II clinical trials

for chronic lymphocytic leukemia therapy (see www.clinicaltrials.gov NCT01419691). Therefore,

auranofin is likely the most ideal candidate to develop for pancreatic cancer therapy that targets

the Trx system to augment oxidative stress.

1.4 A New Therapeutic Strategy for Pancreatic Cancer

The hallmarks of cancer describe essential alterations involved in tumour development. They

include the sustainment of proliferative signaling, evasion of growth suppressors, activation of

invasion and metastasis, enabling replicative immortality, induction of angiogenesis, resisting cell

death, deregulation of cellular energetics, and avoidance of immune destruction263. Redox

signaling and the responses to oxidative stress are involved in every hallmark of cancer264.

Therefore, consideration for cancer promotion by antioxidants should be made when developing

novel anti-cancer therapies. Normal tissues are resistant to antioxidant loss, which may prevent

harmful toxicities of antioxidant inhibition.

As previously discussed, BSO and auranofin appear to be useful drug candidates for combined

targeting of the GSH and Trx systems. Evidence suggests that these systems overlap to compensate

and protect against inefficiencies or inhibition of either system. The Trx system can be an

alternative route for cells to reduce oxidized GSH in vivo265. In addition, physiologic

concentrations of GSH and Grx are capable of reducing oxidized Trx in vitro, suggesting GSH as

a backup to TrxR function266. However, this observation was made in a cell-free system which

may limit its applicability in the complex cell environment.

Page 39: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

29

Combination therapy using both BSO and auranofin can help to overcome compensation. Ideally

the combination of the two drugs will produce improved effectiveness compared to monotherapy

with either compound. This result can either be stated as synergy or potentiation. Synergy is

observed when the effect of two or more drugs is greater than the sum of effects when the drugs

are administered as monotherapies. The effect is mutual in that each drug enhances the effect of

the other267. In contrast, a scenario by which the enhancement of effect is one-sided can be deemed

a potentiation or augmentation. Evaluation of a potentiation can be stated as a fold change in the

effect of one drug, caused by the effect of the other drug267.

Previous studies have shown that dual-inhibition of the GSH and Trx systems can lead to greater

anti-cancer responses127. Use of BSO and auranofin in combination sensitized mouse tumours to

radiotherapy83, and mesothelioma cells show extensive oxidative stress and cell death following

combination of the drugs268. These studies do not evaluate potential synergism of the combination

of BSO and auranofin. Nor do any studies evaluate this combination in pancreatic cancer, which

seems to be an exceptional target of pro-oxidant therapy.

Considering that the Trx and GSH systems comprise the main antioxidant systems in the cell, I

hypothesize that combination therapy using auranofin and BSO will cause primary pancreatic

cancer cells to undergo apoptosis by exacerbating oxidative stress. Auranofin will be identifiable

in pancreatic PDX tumours using IMC. I have three aims to test this hypothesis:

Aim 1: Evaluate the effect of combination therapy on cell viability and death.

Aim 2: Evaluate the effect of combination therapy on the redox environment of the cell.

Aim 3: Evaluate the drug-target interaction of auranofin-TrxR in vitro and biodistribution of

auranofin in vivo.

Page 40: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

30

Chapter 2

Targeting the Thioredoxin and Glutathione Antioxidant Systems to Treat Pancreatic Cancer

Francesco Fazzari1, Samir Barghout2, Qing Chang3, David Hedley1,2,4

1Department of Laboratory Medicine and Pathobiology, University of Toronto, Toronto, Ontario, Canada 2Department of Medical Biophysics, University of Toronto, Toronto, Ontario, Canada 3Fluidigm Canada Inc, Markham, Ontario, Canada 4Division of Medical Oncology and Hematology, Princess Margaret Cancer Centre, Toronto, Ontario, Canada

The work presented in this chapter is presented as a manuscript for future submission.

Page 41: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

31

2.1 Introduction

Pancreatic cancer is the 4th leading cause of cancer death in Canada, however it is only the 12th

most commonly diagnosed2. There is poor prognosis with an overall 5-year survival rate of 5%4.

Significant advancements in the development of new treatment options for the disease have not

been made. Many patients with advanced disease are given standard chemotherapy drugs that

provide little benefit, leading to an urgent need for new strategies to combat this disease.

Pancreatic cancer, among other cancers, is characterized by excessive oxidative stress48, which

occurs when production of intracellular reactive oxygen species (ROS) outpaces the capacity of

antioxidants to protect from ROS. These ROS are produced in mitochondria at an extreme rate in

transformed cells due to inherently high metabolism, and may promote aggressive tumour

features50. Adaptations to protect from ROS are required for cancer cells to survive. These

adaptations primarily involve oxidative stress signaling through Nrf2, which mediates the

oxidative stress response to promote the production of the antioxidants glutathione (GSH) and

thioredoxin (Trx)54. The main oncogenic driver of pancreatic cancer, KRAS, induces Nrf2

activation to maintain antioxidant defenses47. In conjunction to stimulating antioxidant production,

pancreatic cancer relies on glutamine metabolism for the production of NADPH108. This adaptation

occurs because of the need to protect against high ROS during oxidative stress, as NADPH

provides reducing power to antioxidant systems in the cell to prevent cytotoxic damage caused by

high ROS levels.

ROS exhibit dual roles; lower concentrations of ROS that are promoted by antioxidants can

stimulate cancer cell proliferation, whereas higher concentrations of ROS are damaging to

cells49,52. Therefore, antioxidants may contribute to aggressive growth and metastasis in cancer

because of their functions in ROS scavenging, preventing ROS-mediated cellular damage, and

Page 42: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

32

promoting ROS-mediated proliferative signalling148,269. The reliance of pancreatic cancer on

antioxidants to protect from oxidative stress may provide a unique vulnerability in the tumour.

Targeting antioxidants may promote cytotoxic damage specifically in tumours by augmenting

oxidative stress.

Both the GSH and Trx systems have been extensively studied and tested inhibitors exist. GSH is

the most abundant antioxidant cofactor in cells and is responsible for ROS scavenging148. The rate-

limiting step of GSH synthesis is catalyzed by glutamate-cysteine ligase (GCL)148. An inhibitor of

GCL, buthionine sulfoximine (BSO), has been well characterized and previously tested in human

subjects with minimal toxicity197,198. Trx reduces aberrant protein disulfides caused by oxidative

stress269. Oxidized Trx is recycled by thioredoxin reductase (TrxR) using NADPH to replenish its

reducing power256. Three TrxR isozymes exist: cytosolic TrxR1 that has been identified as a fitness

gene209, mitochondrial TrxR2, and testes-specific TrxR3. The gold complex auranofin is an

inhibitor of all three TrxR proteins and has previously been used in the treatment of rheumatoid

arthritis255. It has garnered interest for repurposing as a therapy in various cancers260,261,270.

Many recent studies have identified oxidative stress as an effective target in cancer therapy,

including in KRAS driven tumours75,128,271–273. Targeting one antioxidant pathway may allow for

compensation from another to continue providing protection to the cell204. This may possibly be

overcome by a double hit. To our knowledge no studies have attempted a double hit in pancreatic

cancer. Considering that Trx and GSH comprise the main antioxidant systems in the cell,

combination therapy using auranofin and BSO may cause cell death by depleting antioxidant

capacity and increasing ROS in primary pancreatic cancer cells.

Page 43: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

33

2.2 Materials & Methods

Cell Culture and Reagents

Primary cell lines (9A and OCIP23) were derived from human pancreatic cancer tumours. 9A was

maintained in Roswell Park Memorial Institute (RPMI) 1640 (Wisent) medium supplemented with

10% fetal bovine serum (FBS) (VWR Seradigm Life Science) and 1x penicillin/streptomycin

(Wisent). OCIP23 was maintained in Dulbecco’s Modified Eagle Medium (DMEM)/F-12

(Wisent) supplemented with 2.5% FBS and 1x penicillin/streptomycin. Cells were cultured at 37C

in a humidified incubator with 5% CO2.

Auranofin (Sigma-Aldrich) was diluted in DMSO (Sigma-Aldrich) at 100mM, aliquoted and

stored at -10C to be used once without refreezing. BSO (Sigma-Aldrich) was prepared fresh in

sterilized MilliQ water at 100mM.

MTS Assay

Cell viability was assayed by MTS assay and performed according to the manufacturer’s protocol

(Promega). In summary, 96 well plates were seeded with 5000 cells/100L/well and incubated for

24 hours. Medium was replaced with medium containing DMSO (Sigma-Aldrich) [0.003% for

OCIP23, 0.004% for 9A], auranofin, BSO, or combination of auranofin and BSO at the indicated

concentrations. Each treatment was performed in triplicate. Cells were incubated for 24 hours, then

cell viability was determined by adding 20L of MTS reagent followed by incubation for 3 hours

at 37C and absorbance reading at 492nm using the Multiskan EX microplate spectrophotometer

(Thermo Fisher).

Clonogenic Assay

Page 44: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

34

Clonogenic assays evaluated drug treatment effects on colony formation and was performed as

previously described274. Briefly, cells were grown to 70-80% confluence in 10cm plates then

treated with vehicle (1L/mL DMSO), auranofin, BSO, or a combination at the indicated

concentrations for 24 hours. Cells were harvested and reseeded in 6-well plates at 105, 104, 103,

and 102 cells/well in triplicate, then allowed to incubate for 12 days before fixing and staining with

0.2% methylene blue in 50% ethanol/water. Colonies were counted using the Gel Count (Oxford

Optronix) counter. Plating efficiency was calculated as the average number of colonies divided by

the number of cells plated. Surviving fraction was expressed as the plating efficiency of the sample

divided by the plating efficiency on vehicle control plates.

Multiparametric Flow Cytometry

Apoptosis, ROS levels, and GSH levels were assayed by flow cytometry. 7x105 cells/plate were

seeded in 6cm plate. After 24 (OCIP23) or 48 hours (9A), media was replaced with either normal

medium or BSO-supplemented medium. Auranofin was diluted in DMSO to 1000x the indicated

concentrations and added at 1L/mL. After 24 hour treatment the cells were trypsinized and

resuspended in medium. Cells were centrifuged at 1100rpm for 5 minutes, then the pellets were

resuspended in 0.5mL serum free media. Cell suspensions were then incubated in a water bath at

37C with 40nM DiIC1(5) (ThermoFisher Scientific) and 5M 2’,7’-dichlorofluorescein (DCF)

(ThermoFisher Scientific) for 30 minutes. 5 minutes before the end of the incubation, 40M

monobromobimane (mBBr) (ThermoFisher Scientific) and 1g/mL propidium iodide (PI)

(ThermoFisher Scientific) were added. Following incubation, samples were cooled on ice for 10

minutes then analyzed with the CytoFLEX flow cytometer (Beckman Coulter). During the time

course experiment samples were prepared in the same manner, however the drug treatments were

Page 45: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

35

for 2 and 5 hours, and mBBr was excluded from the panel of fluorescent probes. Data analysis was

performed using FlowJo Software (BD).

Tumour sample preparation

Animal experiments were carried out according to protocols approved by the University Health

Network Animal Care Committee. Patient-derived xenograft (PDX) mice were established

subcutaneously in 5 week old severe combined immunodeficiency (SCID) mice. Two models,

OCIP83 and 9A were used. Mice were given an intraperitoneal injection of 10mg/kg auranofin for

a 24 hour treatment. Mice were later given 120mg/kg pimonidazole (Hypoxyprobe, Inc.), followed

by 30mg/kg EF5 (obtained from Dr. Cameron J. Koch, University of Pennsylvania) 5 and 3 hours

prior to sacrifice, respectively. Tumours were excised, fixed, and processed for paraffin

embedding.

Immunohistochemistry (IHC)

Formalin-fixed, paraffin-embedded (FFPE) sections of subcutaneous PDX tumours were

incubated with mouse pimonidazole antibody (Hypoxyprobe, Inc.) at 1:400, overnight at 4C.

Biotinylated horse anti-mouse IgG was added at 1:200 concentration and incubated for 1 hour.

Detection was performed using the HRP labeling (Vector Labs) and DAB chromogen (Dako).

Imaging Mass Cytometry (IMC) Staining, Data Acquisition, and Visualization

Imaging mass cytometry was used to directly image the distribution of atomic gold in tissue

sections following treatment with auranofin. Tissue staining for IMC was performed as previously

described275. In summary, FFPE sections were dewaxed and rehydrated as routine. Antigen

retrieval and blocking with BSA was performed, followed by incubation with primary antibody

Page 46: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

36

against pimonidazole for 1 hour at room temperature (RT). Samples were then incubated overnight

at 4C with a metal-conjugated antibody cocktail diluted in PBS/0.5% BSA (Table 1). Following

this, samples were washed and exposed to 25M Ir-Intercalator (Fluidigm) for 30 minutes at RT.

Lastly, the samples were rinsed then dried before IMC ablation.

Regions of interest were determined by assessing IHC samples stained for pimonidazole. During

data acquisition, the IMC immunostained samples are inserted into a laser ablation chamber where

the tissue is ablated in a 1m diameter spot. The ablation spot is vaporized, and the plume is carried

with high time-fidelity into the inductively coupled plasma ion source for simultaneous analysis

by the mass cytometer (Fluidigm). The metal isotopes located on each spot are simultaneously

measured. This process continues for the entire area, eventually yielding an intensity map of the

targets throughout the tissue region of interest.

Data-processing and image visualization was performed using the in-house-developed MCD

Viewer (Fluidigm). The signal data from the mass cytometer was exported in text format, then

reconstructed into images. The epithelial cells were defined by E-cadherin and pan-keratin

staining, and the stromal cells were defined by collagen and SMA. Areas within these regions

were selected in the MCD viewer to semi-quantitatively determine the average number of gold

ions within them

Cell Lysis and Protein Determination

Lysis was performed using a lab-made lysis buffer containing 50mM Hepes (pH 8.00), 10%

glycerol, 1% Triton-X, 150mM NaCl, 1mM EDTA, 1.5mM MgCl2, 100mM NaF, 10mM

Na4P2O710H2O, 1mM Na3VO4, and 1 tablet/7mL buffer cOmplete protease inhibitor cocktail

(Sigma-Aldrich). Cell cultures were incubated with 700L lysis buffer for 1 hour at 4C, then

Page 47: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

37

collected and centrifuged at 14,000rpm for 15 minutes at 4C. Supernatants were collected and

stored at -80C until use. Protein determination was performed using the BCA Assay kit

(ThermoFisher Scientific) as per the manufacturer’s protocol.

Gel Electrophoresis and Immunoblot

Samples were resolved using 10% SDS-PAGE. Following transfer to polyvinylidene difluoride

membranes, proteins were detected with antibodies to TrxR1 1:1000 (Cell Signaling Technology),

TrxR2 1:1000 (Cell Signaling Technology), -actin 1:7000 (Abcam), GAPDH 1:4000 (Cell

Signaling Technology). Horseradish peroxidase (HRP) conjugated secondary sheep anti-mouse

(Abcam) and donkey anti-rabbit (Abcam) were used for detection at 1:5000. Quantification was

performed by densitometry using ImageJ (NIH).

Cellular Thermal Shift Assay

Cellular Thermal Shift Assay (CETSA) was performed as previously described276 to assess

auranofin-TrxR engagement. Briefly, to determine the optimal temperature at which maximal

thermal shifts were observed, cells were treated with vehicle or 10M auranofin for 1 h, then

divided into equal portions, centrifuged, and resuspended in PBS with protease inhibitors. Cells

were then heated at temperatures ranging from 46-70C for 3 min in a thermal cycler (SimpliAmp,

Applied Biosystems), cooled at RT for 3 min, snap-frozen in liquid nitrogen, thawed at RT, and

then lysed by 3 freeze/thaw cycles with vortexing in between. Samples were then centrifuged at

20,000 x g for 20 min at 4C and supernatants were collected. Lysates were analyzed by

immunoblotting. When evaluating changes with different auranofin concentrations the same

procedure was followed, however cells were treated with increasing auranofin concentrations at

the determined optimal temperature of 56C.

Page 48: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

38

Statistical Analysis

Data are presented as the mean value S.E.M. GraphPad Prism version 7.0 (GraphPad Software,

Inc., San Diego, CA, USA) was used for statistical analysis. A value of P<0.05 was considered to

indicate statistically significant differences.

2.3 Results

Auranofin, but not BSO, inhibits cell growth and viability of primary pancreatic cancer cell

lines.

To evaluate the effect of Auranofin and BSO on cell viability, human primary pancreatic cancer

cells were treated with different concentrations of either drug for 24 hours. MTS assay measured

cell metabolic activity. BSO alone displayed no significant effect on the viability of the primary

cells (Fig. 2.1A). In contrast, auranofin reduced viability in a concentration dependent manner with

IC50 values 1.40μM and 2.35μM in OCIP23 and 9A, respectively (Table 2.1; Fig. 2.1B). These

data suggest that after 24 hours, these two primary pancreatic cancer cell lines are susceptible to

auranofin, but not BSO. Since BSO does not reduce viability at this time point, any interaction

between the two drugs cannot be deemed synergy.

BSO potentiates auranofin’s inhibitory effect on cell viability of primary pancreatic cancer cells.

Using the MTS assay, we evaluated cell viability following treatment of cells with a combination

of auranofin and BSO. Co-treatment with 10μM BSO caused a 61.1-fold decrease in the IC50 of

auranofin in OCIP23 (IC50 = 22.9nM), and a 4.07-fold decrease in 9A (IC50 = 578nM) (Fig.

2.1B). The addition of 50μM BSO caused a 238-fold decrease in the IC50 of auranofin in OCIP23

(IC50 = 5.88nM), and a 11.0-fold decrease in the IC50 in 9A (IC50 = 213nM) (Fig. 2.1B). These

Page 49: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

39

data support a strong BSO-mediated potentiation of auranofin cytotoxicity. The higher IC50 and

lower potentiation observed with 9A cells suggest they possess a higher reserve capacity of

TrxR1/2 activity that makes them more resistant to drug-induced oxidative stress. To test this

hypothesis, we assessed expression levels of Trx1/2 in 9A and OCIP23 cells. 9A cells showed a

significantly higher expression level of TrxR1, but not TrxR2, compared to OCIP23 cells

(p=0.0007) (Fig. 2.1C and D). This may explain, at least in part, the differential response of

OCIP23 and 9A to auranofin and combination with BSO.

Combination of BSO and auranofin inhibits clonogenicity of primary pancreatic cancer cells.

We then conducted clonogenic assay to determine whether the BSO-auranofin combination

inhibited long-term viability and if low auranofin concentrations on their own improved viability

as observed by MTS assay. There was a significant decrease in clonogenicity following

combination treatment in 9A (p=0.0155) with 43.1% surviving fraction (Fig. 2.1E) and OCIP23

(p=0.0088) with 43.8% surviving fraction (Fig. 2.1F) when compared to vehicle controls. Single

treatment with auranofin (100nM or 400nM) showed no increase in clonogenicity, which contrasts

the improved viability of low auranofin concentrations observed via MTS assay (Fig. 2.1B). The

results of these two assays may differ because of the different endpoints measured, metabolism

versus colony growth. These data support the potentiation of auranofin by BSO at low auranofin

concentrations that otherwise have no effect on cell death and viability when used as monotherapy.

Table 2.1. IC50 values of auranofin monotherapy and in combination with BSO in primary

pancreatic cancer cell lines.

Cell Line BSO Concentration (M) Auranofin IC50 (nM)

OCIP23 - 1400

OCIP23 10 22.9

OCIP23 50 5.88

9A - 2350

9A 10 578

9A 50 213

Page 50: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

40

Figure 2.1. Dose-response of antioxidant pathway inhibitors, TrxR isozyme levels, and

colony formation potential. (A and B) Viability of primary pancreatic cancer cells was

determined after 24 hour treatment using an MTS assay. Shown are the toxicity profiles of A) BSO

monotherapy (n=3), and B) auranofin monotherapy or combination (n=3). (C and D) Relative

amounts of C) TrxR1 and TrxR2 in untreated cell lines (n=3) and their corresponding D)

immunoblots. Levels were normalized to the first OCIP23 sample. (E and F) Colony formation of

primary cell lines was determined after 24 hour treatment. Cells were reseeded and incubated for

12 days before colonies were assesed for E) 9A (n=3) (Combination of Au 400nM and BSO 10M)

and F) OCIP23 (n=3) (Combination of Au 100nM and BSO 10M). Combinations were compared

to vehicle controls. Surviving fraction is represented as a percentage of vehicle controls. Shown

are the S.E.M. Significant decrease from vehicle control, * = p<0.05, ** = p<0.01, *** = p<0.001.

0 20 40 60 80 100 1200

20

40

60

80

100

120

BSO concentration (uM)

% C

ell V

iabil

ity

BSO Dose-Response

OCIP239A

9A - 10uM BSO9A - 0 BSO

9A - 50uM BSO

OCIP23 - 50uM BSOOCIP23 - 10uM BSO

10-3 10-2 10-1 100 101 102 103 104 1050

20

40

60

80

100

120

140

160

180

200

220

Auranofin concentration (nM)

% C

ell V

iabil

ity

Auranofin Dose-Response with and without BSO

OCIP23 - 0 BSO

Vehic

le

BSO

10u

M

Au

400n

M

Au

100n

M

Com

bina

tion

0

20

40

60

80

100

120

140

Surv

ivin

g F

ract

ion (

%)

OCIP23

**

9A OCIP23

TrxR1

β-actin

TrxR2

β-actin

A) B)

C)

E) F)

D)

Page 51: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

41

BSO-auranofin combination induces apoptosis in primary pancreatic cancer cells.

We conducted live cell flow cytometry to determine whether apoptosis contributed to cell death

after BSO-auranofin combination. The loss of mitochondrial membrane potential (MMP) has been

observed to be one of the first steps of apoptosis, and has previously been observed following

auranofin treatment258. In addition, the loss of membrane integrity, and subsequent uptake of

propidium iodide (PI), is a marker of cell death. Here we used DiIC1(5), a marker of MMP, and

PI to assess cell death after combination treatment. Auranofin or BSO alone did not increase the

MMP negative or PI positive populations. In contrast, combinations of these low concentrations

increased the proportion of cells observed to be starting apoptosis and those that have already died

in both OCIP23 (Fig. 2.2) and in 9A. These results indicate that combinations of low

concentrations of auranofin and BSO induces apoptosis in primary pancreatic cancer cells.

BSO-auranofin combination reduces thiol levels and induces an acute increase in ROS levels.

We also used live cell flow cytometry to assess the effect of combinatorial treatment on oxidative

stress. First, GSH was examined after 24 h treatment by using mBBr, which fluoresces after

binding to free thiols. GSH is the most prominent free thiol in cells, and contributes to

approximately 50% of the mBBr signal277. The treated cells were gated for PI negative populations

to restrict analysis to viable cells. Auranofin did not affect GSH levels when applied as a

monotherapy in OCIP23 but had a significant impact in 9A (p=0.0345). However, BSO had a

significant impact (OCIP23 p<0.0001; 9A p=0.0003) on GSH levels (Fig. 2.4A). The GSH status

of MMP negative populations, both PI negative and positive, was assessed during the same live

cell flow cytometry assay. Following 24 hour treatment both populations displayed depleted GSH

compared to MMP positive populations (Fig. 2.3). These data suggest that the combination therapy

may lead to cytotoxicity by depletion of GSH levels.

Page 52: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

42

Figure 2.2. Representative two parameter plots measuring cell death and apoptosis in

OCIP23. Cells were treated with either monotherapy or combination therapy for 24 hours

followed by flow cytometry. Apoptotic cell populations are MMP-, PI- (bottom left quadrant);

Dead cell populations are MMP-, PI+ (Upper left quadrant); Live cells are MMP+, PI- (bottom

right quadrant). Shown are the percentages of the total population for each subpopulation.

Auranofin concentration

0nM 6nM 30nM 75nM

BS

O co

ncen

tration

M

10

μM

5

M

MMP

PI

Page 53: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

43

Figure 2.3. Representative two parameter plots assessing the GSH content of apoptotic cells

in OCIP23. Cells were treated with either monotherapy or combination therapy for 24 hours

followed by flow cytometry. MMP is a marker for apoptosis, mBBr is a marker for GSH, and PI

is a marker for cell viability. PI positive (non-viable) cells are shown in red, and PI negative

(viable) cells are shown in grey.

Second, ROS were examined following 2 and 5 hour treatments by 2’,7’-dichlorofluorescin (DCF),

which fluoresces following oxidation by ROS. Again, treated cells were gated for PI negative

populations to restrict analysis to viable cells. At these time points, ROS were not affected by

auranofin alone. However, BSO alone and in combination with auranofin displayed a significant

dose-dependent increase (OCIP23 p<0.0001; 9A p<0.0001) in ROS levels (Fig. 2.4B). Combined,

these results indicate that the combination of auranofin and BSO leads to oxidative stress in the

Auranofin concentration

0nM 6nM 30nM 75nM

BS

O co

ncen

tration

M

10

μM

5

M

mBBr

MM

P

Page 54: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

44

treated cells. This suggests that depletion of GSH and increase in ROS may contribute to cell death

after combination therapy.

Figure 2.4. Effect of treatment on thiol and ROS levels in 9A and OCIP23 cell lines. A) Mean

fluorescence intensity (MFI) of mBBr in PI negative cells after 24 hour treatment with BSO,

auranofin, or combination (9A n=3; OCIP23 n=3). B) MFI of DCF in PI negative cells after 2 and

5 hour treatment with BSO, auranofin, or combination (9A n=3; OCIP23 n=3). Fluorescence was

normalized to non-treated controls. Shown are the S.E.M.

Auranofin binds to TrxR1 in primary pancreatic cancer cells at concentrations associated with

oxidative stress and cell death.

We performed a cellular thermal shift assay (CETSA) to observe whether auranofin engages its

targets TrxR1 and TrxR2 in intact primary pancreatic cancer cells at different concentrations (Fig.

B) A)

Page 55: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

45

2.5). This assay is based on thermostabilization of protein targets upon binding chemical

ligands276. It is more clinically relevant as it allows to test drug-target engagement in the complex

cellular environment, as opposed to cell-free assays. As an initial step in CETSA, we treated 9A

cells with a saturating concentration (10 µM) of auranofin and determined the optimal temperature

at which we observed maximal thermostabilization of TrxR1/2 in treated cells compared to

untreated controls. By analysis of thermal shift blots, we found 56C to be the optimal temperature

for TrxR1 and TrxR2 (Fig. 2.5A).

We then tested auranofin engagement after treating 9A and OCIP23 cells with increasing

concentrations to determine whether it binds TrxR1/2 at concentrations associated with cell death.

In both 9A and OCIP23 we observed TrxR1-auranofin binding at concentrations as low as 78nM

(Fig. 2.5B and C). Binding was observed at concentrations much below the IC50’s of auranofin

monotherapy in both cell lines. In addition, our CETSA data suggest that auranofin-mediated cell

death is attributable to targeting TrxR1, but not TrxR2, at IC50’s observed after BSO combination

(Table 2.1). Since OCIP23 and 9A cells exhibit different TrxR1 but not TrxR2 expression levels,

our CETSA data provide further explanation of greater potentiation observed with OCIP23

compared to 9A cells. Taken together, our data show that auranofin bound TrxR1 in intact primary

pancreatic cancer cells at concentrations capable of inducing cell death, and auranofin targets

TrxR1 not TrxR2 at concentrations associated with BSO-mediated potentiation.

Page 56: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

46

Figure 2.5. Auranofin-TrxR interaction in vitro. A) Engagement of auranofin and TrxR was

assessed by CETSA. 9A was treated with 10M auranofin for 1 hour. Following drug treatment,

cells were exposed to the indicated temperatures for 3 minutes prior to lysis and immunoblot (n=1).

B) 9A and C) OCIP23 were treated with different auranofin concentrations for 2 hours before heat

treatment at 56C and immunoblot (n=1).

GAPDH

TXRX2

TXRX1

UT T

46°C

UT T

49°C

UT T

52°C

UT T

55°C

UT T

58°C

UT T

61°C

UT T

64°C

UT T

67°C

UT T

70°C

Auranofin (µM)

TRXR1

GAPDH

TRXR2

Auranofin (µM)

TRXR1

GAPDH

TRXR2

A)

B)

C)

Page 57: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

47

Table 2.2. Panel of markers measured by IMC using subcutaneous PDX tumours.

Antibody/Reagent Antibody Clone Metal Final Concentration

(g/mL)

Avanti Lipid N/A 115In 10M

-SMA 1A4 141Pr 1.25

Thioredoxin 2G11/TRX 146Nd 10

EF5 ELK3-51 149Sm 10

Pimonidazole N/A 155Gd 10

E-cadherin 24E10 158Gd 2.5

Pan-keratin C11 164Dy 0.6

Goat Anti-Mouse

(Erg1 detection)

N/A 165Ho 1:100

Collagen Goat polyclonal 169Tm 1.25

Intercalator N/A 191/193Ir 0.25M

Auranofin N/A 197Au 10mg/kg

Auranofin distributes within PDX tumours 24 hours after treatment, and associates with the

stroma.

Auranofin contains gold in its molecular structure, which suggested the possibility to detect it in

tissues using imaging mass cytometry (IMC), as we have recently described for platinum-based

drugs275. This novel high throughput technology detects heavy metal isotopes on a histological

section. These heavy metals can either be linked to antibodies or found within the tissue. Data

gained from this method can give information on the distribution of auranofin, its interactions with

the tumour tissue, and an approximation of its concentration within different compartments of a

tumour as long as auranofin is detectable with IMC. The antibody panel used can be seen in Table

2.2. In this pilot study we observed auranofin distributed throughout the two PDX tumours (Fig.

2.6). Auranofin appeared to be more closely associated with collagen in the stromal compartment

of the tumour (Fig. 2.6A and C), rather than E-cadherin in the epithelial compartment (Fig. 2.6B

and D). We then approximated the average number of gold ions per pixel (each pixel has a defined

volume of 5m3) in each compartment (Table 2.3). This semi-quantitative method showed a

Page 58: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

48

preferential accumulation of auranofin in the stroma. This method can be expanded in the future

to gain a greater understanding of auranofin’s effects in vivo.

Figure 2.6. Distribution of auranofin in subcutaneous PDX tumours. Mice were treated with

10mg/kg auranofin for 24 hours, followed by pimonidazole for 5hr and EF5 for 3 hours prior to

sacrifice. Tissue section ablations were performed using IMC and markers were false-coloured.

The relation of auranofin (green) to (A and C) stromal collagen (blue), (B and D) epithelial E-

cadherin (blue), and pimonidazole (red) in OCIP83 (A and B) and 9A (C and D) tumours is shown.

Scale bars are provided for reference. Arrows highlight regions of interest.

OC

IP83

9A

A B

C

200μm

D

200μm 200μm

200μm

Page 59: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

49

Table 2.3. Estimation of the mean ions/pixel in the stroma and epithelium of subcutaneous

tumours from auranofin treated PDX mice.

Model Ions/pixel

(Stroma)

Ions/pixel

(Epithelium)

OCIP83 4.75 2.83

9A 8.67 8.28

2.4 Discussion

New strategies are required in the treatment of pancreatic cancer. Oxidative stress is an inherent

feature in pancreatic cancer that can be potentially exploited by therapy. The described results

provide evidence that a combination of BSO and auranofin at low clinically achievable drug

concentrations278,279 yields anticancer efficacy in vitro by augmenting oxidative stress.

In this study we assessed the effect of combination therapy on cell death and the redox environment

of the cell. Through an MTS assay and flow cytometry-based assays we identified a strong

potentiation of the cytotoxic effects of auranofin by BSO. Although BSO did not contribute to cell

death on its own after 24 hour treatment, it was responsible for causing increased ROS and

decreased GSH during the combination therapy. Low auranofin concentrations did not

independently affect ROS and had a minimal effect on GSH in one cell line. However, it produced

cytotoxic effects when combined with BSO. Overall, these findings are consistent with the

predicted mechanism of action of a decreased antioxidant capacity and increased oxidative stress

following combination of two antioxidant inhibitors64,83,204. However, it was surprising to observe

that low auranofin concentrations did not independently affect ROS. Other studies have assessed

ROS following treatment with higher auranofin concentrations, and observed substantial

increases280. Our results suggest redundancy between the antioxidant pathways in that both GSH

Page 60: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

50

and Trx moderate ROS levels, however the low auranofin concentrations are not required to affect

ROS because of the effect of BSO in this context. It is possible that only a modest increase in ROS

is required to kill the cells because of the concomitant depletion of antioxidants, although to our

knowledge literature does not exist that elucidates the amount of ROS required for cytotoxicity in

the context of antioxidant depletion. Moreover, some cytotoxicity can likely be attributed to

downstream action of TrxR. As protein disulfides accumulate under oxidative stress in the cell,

Trx is unable to remedy the damage because of impaired recycling through TrxR. Cells would lose

their ability to maintain critical functions. These data intimate that these two antioxidant systems

are critical to cancer cell survival. It is possible that other inhibitors towards these systems can be

used to achieve similar results.

When assessing the engagement of auranofin and TrxR we utilized CETSA. This assay provided

the capability to determine whether auranofin bound its target in the context of a complex

biological system276. Our observations show that auranofin engages TrxR1 at low concentrations

compatible with cell death. Engagement with TrxR2 was not observed in this assay. This data is

significant as it shows that auranofin interacts with its target in intact cells, whereas results

obtained in cell-free systems eliminate the presence of other proteins that may interfere with this

binding. This result provides greater support for auranofin’s proposed target, TrxR1, mediating the

observed cytotoxicity. Future studies to show auranofin-TrxR1 engagement in vivo can further

support this finding.

To assess the biodistribution of auranofin in the tumour we used IMC. This provided the unique

insight into auranofin’s distribution 24 hours following treatment. In this pilot study we observed

auranofin associated with collagen in the stromal compartment of the tumour. Although this has

not been previously reported, IMC has previously observed cisplatin association with collagen in

Page 61: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

51

PDX tumours275. This finding is corroborated by the long half-life that has been observed in

patients taking oral auranofin281. Collagen may serve as a reservoir for the drug and contribute to

its slow elimination. Our semi-quantitative approach to estimate the number of ions in each

compartment trended towards an increase in the stroma compared to the epithelium, although the

data was not significant. In the future this method can be applied to gaining an in depth

understanding of the interaction of auranofin and molecular markers of oxidative stress to observe

how auranofin affects the molecular biology of the tumour.

This study is limited in a few different aspects. First, the measure of GSH by mBBr fluorescence

using flow cytometry is not specific for GSH. mBBr binds to all free thiols in cells. However, the

background we captured is consistent with previously reported results277. Second, although our

data shows that there is an interaction of auranofin and TrxR1 in the cells, there is no functional

output to measure the effect on TrxR1 activity. Therefore, it is difficult to determine how effective

the low auranofin concentrations are at inhibiting TrxR1 activity. Third, the semi-quantitative

method of determining the number of ions in the tumour can only provide an estimate, as only

small regions of stroma and epithelium in each section could be analyzed, and the number of ions

lost during acquisition cannot be accounted for.

In conclusion, we propose that targeting antioxidants can prove to be an effective new strategy in

treating pancreatic cancer. In the context of the tested combination of BSO and auranofin, we

observed that BSO potentiates auranofin by augmenting oxidative stress, allowing auranofin to

elicit its cytotoxic effects at lower concentrations. Future in vivo studies are required to assess

whether this method is feasible in the context of a living organism, as it is possible that the double

hit may cause undesired toxicities in normal tissues. However, if basal oxidative stress occurs in

the tumour then the reliance on antioxidants should make them more vulnerable to antioxidant

Page 62: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

52

inhibition. Furthermore, normal tissues, including hepatocytes, can withstand oxidative stress

following Trx and GSH loss with dietary methionine253. This work gives further evidence to

support the growing body of evidence that suggests that cancers rely on antioxidants for their

survival. Hopefully, this can be a step towards the future development of effective treatments for

pancreatic cancer.

Page 63: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

53

Chapter 3

Overall Discussion, Future Directions, and Conclusions

Page 64: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

54

3.1 Overall Discussion

Oxidative stress is an interesting target for future cancer therapy development. Pancreatic cancer

relies on antioxidants to protect from oxidative stress and promote proliferation. As such, it is

sensitive to antioxidant depletion which may alter the redox environment. The reliance on

antioxidants is specific to tumour tissue, whereas normal tissues may resist antioxidant loss253.

Targeting antioxidant systems may require a double hit as they can compensate for each other’s

functions.

A common barrier to the development of novel therapies is the length of time from drug discovery

to implementation in clinics282. However, the strategy of augmenting oxidative stress avoids this

barrier with the use of auranofin, which has been used in clinics, and BSO, which has been used

in human clinical trials. Before a combination of the two can be used in human trials, preliminary

tests are required to examine their effects in vitro and in vivo. Therefore, the goal of this thesis is

to test the combination against pancreatic cancer and identify how the drugs interact.

3.1.1 BSO potentiates the acute cytotoxicity of auranofin

I first sought to examine the dose-response of primary human pancreatic cancer cells to both BSO

and auranofin. I showed that BSO potentiates auranofin’s cytotoxicity at 24 hours and that

response may depend partially on TrxR1 levels. Previous studies have examined the effect of these

drugs on cell death, in particular apoptosis. At 24 hours BSO does not induce apoptotic cell

death149. In contrast, treatment of less than 24 hours with auranofin is sufficient to induce the

mitochondrial permeability transition indicative of apoptosis258. Of the TrxR isoforms present in

the cell, TrxR1 inhibition is likely most responsible for mediating the response to auranofin, as it

has been identified as a fitness gene209.

Page 65: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

55

Following 24 hour treatment BSO did not affect viability of either cell line, indicating that the cells

were resistant to acute GSH loss. Over the same timespan auranofin reduced viability in a dose-

dependent manner. OCIP23 was more sensitive than the 9A cell line. Since BSO was not lethal on

its own after 24 hours, synergy was excluded because by definition synergy requires an

independent effect of each drug. I then treated cells with both drugs simultaneously for 24 hours

and observed a strong potentiation of auranofin’s cytotoxic effects by BSO. Again, OCIP23 was

more sensitive. The required concentrations of both drugs was within the safe plasma

concentrations that have been achieved during human clinical trials of either drug278,279. This

supports the idea that these cytotoxic concentrations can be attained if administered to a human

patient. Although the discrepancy in sensitivity between the two cell lines can result from

numerous factors, I observed a significant decrease in TrxR1, but not TrxR2, levels in OCIP23.

As TrxR1 has been identified as a fitness gene this difference may contribute to sensitivity to this

combination therapy and may potentially be validated in the future as a biomarker to predict

response to auranofin therapy.

The MTS data showed an odd result in that low auranofin concentrations seemed to improve

viability. However, this may be an artifact of the MTS method itself, in that it measures NADH

production from mitochondrial metabolism which may accumulate in dysfunctional mitochondria

during oxidative stress283. The clonogenic assay showed that these concentrations did not improve

long-term cell growth, and when used in the combination therapy effectively inhibited

clonogenicity.

3.1.2 Apoptosis contributes to cell death

I also identified that apoptosis contributes to the mechanism of cell death of the combination

therapy. Previous work has shown apoptosis caused by BSO following treatment of 48 hours and

Page 66: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

56

longer149, and apoptosis following short-term auranofin treatment258. The combination also

displayed apoptosis as shown by the mitochondrial permeability transition. However, this cannot

exclude other mechanisms of cell death as contributors, including senescence and ferroptosis.

Ferroptosis in particular may likely contribute as a result of augmented oxidative stress but was

not assessed using the current assays284.

3.1.3 The drugs engage their expected targets

An important observation when assessing a therapy is to determine whether the drugs enact their

expected primary effect. In the case of BSO, GCL inhibition is expected to decrease GSH levels

as previously observed83. I identified that in the primary pancreatic cancer cells there was a

decrease in GSH following both BSO monotherapy and in combination with auranofin. Thus, BSO

enacted its primary role of preventing GSH synthesis. Auranofin is expected to decrease TrxR

activity through direct interactions with TrxR. Although this thesis does not include an assay for

TrxR activity, it is well-established that auranofin can reduce or eliminate TrxR activity83,254.

Instead I used CETSA to examine drug-target interaction in the context of the complex cellular

environment. I observed that auranofin engaged TrxR1 at low concentrations compatible with cell

death during combination therapy, indicating that auranofin does in fact bind its proposed target

in the cell when used at low concentrations.

3.1.4 Oxidative stress follows combination therapy

The downstream effects following target inhibition are important to understand when assessing a

potential new therapy. I observed that simultaneous inhibition of GSH and Trx augments oxidative

stress by increasing ROS. Previous studies have observed this phenomena following both BSO

and auranofin monotherapy83,149,254. As such, the studied combination therapy caused the expected

downstream effect. Of note, the increase in ROS was modest and of lower magnitude than

Page 67: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

57

previously reported with higher auranofin concentrations83,149. However, the observed increase in

ROS was still capable of inducing cell death. Unfortunately, it is difficult to make comparisons to

other work. This is because other studies mostly focus on the amount of exogenous ROS required

to kill cells, rather than ROS from endogenous sources285. Furthermore, the amount of ROS

required to kill in the context of a cell with compromised antioxidant defences has not been

previously determined. It is likely that antioxidant depletion allows for this modest acute spike in

ROS to elicit cytotoxic effects. Lastly, it should be noted that higher auranofin concentrations (ie.

monotherapy IC50 concentrations) have been shown to induce ROS83, which would resolve their

cytotoxic effects as a monotherapy in light of my results showing that low auranofin concentrations

do not independently augment ROS.

3.1.5 IMC is an effective tool to examine auranofin in vivo

Powerful analytical tools are required to understand the full effect of a therapy in vivo. Auranofin’s

unique chemistry provides the potential of IMC identification in tissues, as has been reported with

cisplatin275. My pilot study determined that gold from auranofin can be identified by IMC 24 hours

after drug administration. As with cisplatin, it displays a preferential association with collagen in

the PDX tissue stroma versus E-cadherin in the epithelium275. However, this data does not provide

insight to the biodistribution at shorter time points and may be suggestive of the auranofin reservoir

following elimination of most auranofin molecules. This would be consistent with auranofin’s

exceptionally long total body elimination (55-80 days) observed during clinical trials286. A larger

number of samples are required to prove these results, as this was only a pilot study. Lastly, it

should be noted that IMC cannot distinguish between auranofin and atomic gold. Therefore, the

observations of auranofin distribution in the stroma may be indicative of gold following auranofin

Page 68: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

58

breakdown. However, this is consistent with previous methods that restrict detection to atomic

gold when evaluating auranofin287.

The significance of this finding is that IMC can be a tool for the assessment of a larger cohort to

develop an in-depth study of auranofin’s effects in mouse and/or human clinical trials. An antibody

panel can be developed to examine auranofin’s interactions with different tissue structural features,

cell proliferation or growth arrest, hypoxia, cell signaling, and many other potential features of

interest.

3.2 Future Directions

A number of different avenues can be explored following the work described in this thesis. The

first possibility would be to study the mechanism of action greater detail. As previously mentioned,

apoptosis contributes to cell death following the combination therapy. However, the presented data

cannot exclude other mechanisms of death such as ferroptosis, which can result from excessive

lipid peroxidation that occurs during oxidative stress284. The fluorescent probe C11-BODIPY can

be used to assess the extent of lipid peroxidation in a convenient flow cytometry assay that includes

probes for ROS, GSH, viability, and MMP288. Another avenue could be to perform experiments to

further support the current data. The current method for GSH measurement uses mBBr

fluorescence, which also detects protein thiols. A modified version of the Tietze enzymatic

recycling assay can provide a direct biochemical measurement of GSH following BSO

treatments289. The current method to assess auranofin efficacy detects whether auranofin binds its

target. However, this does not specify the extent of TrxR activity loss. An insulin reduction assay

can be performed to assess TrxR activity at the same concentrations290,291.

Page 69: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

59

To optimize therapeutic results alterations to the dosing regimen can be tested. For example, BSO

concentrations can be decreased while increasing auranofin. This can be changed to ensure that

doses of each drug that are easiest to achieve in humans are used. Furthermore, alternative

treatment lengths can be tested. This may maximize cytotoxic effects of BSO that have been

observed when used for 48 hours and longer149. Lastly, the strategy for therapy can be altered. The

current study uses simultaneous administration of BSO and auranofin. However, since GSH

depletion by BSO requires time to commence it is possible that an alternative strategy may take

advantage of this. BSO pre-treatment for 24 hours may possibly maximize the effect of auranofin

as the second hit would be administered in a GSH knockdown system. Other alterations to the

strategy may include additions to the current combination. For example, lower doses of BSO and

auranofin may sensitize tumours to chemo or radiotherapy83,202. Chemotherapy drugs like

gemcitabine or ionizing radiation can be used in combination with reduced BSO and auranofin

doses to optimize responses.

Lastly, the most significant option for future studies would be to examine the efficacy of this

combination therapy in orthotopic PDX mice. This approach would allow for testing in a model

that is the most representative of tumours seen in clinics. First, a few mice can be used to determine

the safety of the combination, as adverse events can be unpredictable. If the therapy is deemed

safe a larger experiment can be performed. The design of such an experiment would place mice in

4 treatment arms: control, auranofin, BSO, and combination therapy. Numerous endpoints can be

assessed including survival, changes in tumour volume, and number of metastases. In addition, a

comparison can be made to the current standard of care if another group is used. It is possible that

sensitivity to therapy may depend on TrxR1 levels. As such, tumours bearing different amounts of

TrxR1 can be tested.

Page 70: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

60

This experiment also provides the opportunity to assess the effects of auranofin using IMC.

Various questions can be postulated and answered if the correct antibody panel is used on the

tumour tissue. This can also compare differences between monotherapy and combination therapy.

In conjunction, a shorter time-course experiment can examine the timing of different effects

following therapy. These experiments would provide unique insights into how auranofin affects

tumour tissue in vivo, and whether this combination therapy is safe and effective enough to receive

attention in human clinical trials. An additional benefit of running a trial like this would be that

CETSA can be performed on tumours to assess auranofin engagement with TrxR in vivo.

3.3 Conclusions

Oxidative stress is a hallmark of pancreatic cancer that deserves consideration in the development

of new therapies against pancreatic cancer. I conclude in chapter 2 that the current study supports

the efficacy of a combination therapy using BSO and auranofin to treat pancreatic cancer. In this

combination BSO potentiates auranofin’s cytotoxic effects. This therapy results in both antioxidant

depletion and augmentation of oxidative stress. Furthermore, auranofin is detectable by IMC

analysis. Future work can further characterize the effects of this combination and identify whether

it is feasible for clinical use in human pancreatic cancer patients.

Page 71: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

61

References

1. Hidalgo, M. Pancreatic Cancer. N Engl J Med 362, 1605–1617 (2010).

2. Canadian Cancer Society’s Advisory Committee on Cancer Statistics. Canadian Cancer

Statistics 2017. Toronto, ON: Canadian Cancer Society; 2017. Available at:

cancer.ca/Canadian-Cancer-Statistics-2017-EN.pdf (accessed April 2018).

3. Siegel, R. L., Miller, K. D. & Jemal, A. Cancer statistics, 2018. CA Cancer J Clin 68, 7–30

(2018).

4. Garrido-Laguna, I. & Hidalgo, M. Pancreatic cancer: from state-of-the-art treatments to

promising novel therapies. Nat Rev Clin Oncol 12, 319–334 (2015).

5. Flook, R. & van Zanten, S. V. Pancreatic cancer in Canada: incidence and mortality trends

from 1992 to 2005. Can J Gastroenterol 23, 546–550 (2009).

6. Rahib, L. et al. Projecting cancer incidence and deaths to 2030: The unexpected burden of

thyroid, liver, and pancreas cancers in the United States. Cancer Res 74, 2913–2921 (2014).

7. Neoptolemos, J. P. et al. Therapeutic developments in pancreatic cancer: current and future

perspectives. Nat Rev Gastroenterol Hepatol 15, 333-348 (2018).

8. Arbyn, M. et al. European guidelines for quality assurance in cervical cancer screening.

second edition—summary document. Ann Oncol 21, 448–458 (2010).

9. Piccart-Gebhart, M. J. et al. Trastuzumab after adjuvant chemotherapy in HER2-positive

breast cancer. N Engl J Med 353, 1659–1672 (2005).

10. Romond, E. H. et al. Trastuzumab plus adjuvant chemotherapy for operable HER2-positive

breast cancer. N Engl J Med 353, 1673–1684 (2005).

11. Rhim, A. D. et al. EMT and dissemination precede pancreatic tumor formation. Cell 148,

349–361 (2012).

12. Moertel, C. G. Chemotherapy of gastrointestinal cancer. N Engl J Med 299, 1049–1052

(1978).

13. Burris, H. A. et al. Improvements in survival and clinical benefit with gemcitabine as first-

line therapy for patients with advanced pancreas cancer: a randomized trial. J Clin Oncol

15, 2403–2413 (1997).

14. Conroy, T. et al. FOLFIRINOX versus gemcitabine for metastatic pancreatic cancer. N

Engl J Med 364, 1817–1825 (2011).

15. Von Hoff, D. D. et al. Increased survival in pancreatic cancer with nab-paclitaxel plus

gemcitabine. N Engl J Med 369, 1691–1703 (2013).

Page 72: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

62

16. El Rassy, E., Assi, T., El Karak, F., Ghosn, M. & Kattan, J. Could the combination of Nab-

paclitaxel plus gemcitabine salvage metastatic pancreatic adenocarcinoma after

FOLFIRINOX failure? A single institutional retrospective analysis. Clin Res Hepatol

Gastroenterol 41, e26–e28 (2017).

17. Muranaka, T. et al. Comparison of efficacy and toxicity of FOLFIRINOX and gemcitabine

with nab-paclitaxel in unresectable pancreatic cancer. J Gastrointest Oncol 8, 566–571

(2017).

18. Enzler, T. & Bates, S. Clinical trials in pancreatic cancer: A long slog. The Oncologist 22,

1424–1426 (2017).

19. Bagherabad, M. B., Afzaljavan, F., ShahidSales, S., Hassanian, S. M. & Avan, A. Targeted

therapies in pancreatic cancer: promises and failures. J Cell Biochem (2017) doi:

[10.1002/jcb.26284].

20. Kindler, H. L. et al. Gemcitabine plus bevacizumab compared with gemcitabine plus

placebo in patients with advanced pancreatic cancer: phase III trial of the Cancer and

Leukemia Group B (CALGB 80303). J Clin Oncol 28, 3617–3622 (2010).

21. Philip, P. A. et al. Phase III study comparing gemcitabine plus cetuximab versus

gemcitabine in patients with advanced pancreatic adenocarcinoma: Southwest Oncology

Group-directed intergroup trial S0205. J Clin Oncol 28, 3605–3610 (2010).

22. Moore, M. J. et al. Erlotinib plus gemcitabine compared with gemcitabine alone in patients

with advanced pancreatic cancer: a phase III trial of the National Cancer Institute of Canada

Clinical Trials Group. J Clin Oncol 25, 1960–1966 (2007).

23. Kleeff, J. & Michl, P. Targeted therapy of pancreatic cancer: biomarkers are needed. Lancet

Oncol 18, 421–422 (2017).

24. Khalil, D. N., Smith, E. L., Brentjens, R. J. & Wolchok, J. D. The future of cancer

treatment: immunomodulation, CARs and combination immunotherapy. Nat Rev Clin

Oncol 13, 273–290 (2016).

25. Sharma, P., Wagner, K., Wolchok, J. D. & Allison, J. P. Novel cancer immunotherapy

agents with survival benefit: recent successes and next steps. Nat Rev Cancer 11, 805–812

(2011).

26. Hodi, F. S. et al. Improved survival with ipilimumab in patients with metastatic melanoma.

N Engl J Med 363, 711–723 (2010).

27. Robert, C. et al. Pembrolizumab versus ipilimumab in advanced melanoma. N Engl J Med

372, 2521–2532 (2015).

28. Silva, I. P. & Long, G. V. Systemic therapy in advanced melanoma: integrating targeted

therapy and immunotherapy into clinical practice. Curr Opin Oncol 29, 484-492 (2017).

Page 73: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

63

29. Royal, R. E. et al. Phase 2 trial of single agent ipilimumab (anti-CTLA-4) for locally

advanced or metastatic pancreatic adenocarcinoma. J Immunother 33, 828-833 (2010).

30. Brahmer, J. R. et al. Safety and activity of anti–PD-L1 antibody in patients with advanced

cancer. N Engl J Med 366, 2455–2465 (2012).

31. Manji, G. A., Olive, K. P., Saenger, Y. M. & Oberstein, P. Current and emerging therapies

in metastatic pancreatic cancer. Clin Cancer Res 23, 1670–1678 (2017).

32. Brunet, L. R., Hagemann, T., Gaya, A., Mudan, S. & Marabelle, A. Have lessons from past

failures brought us closer to the success of immunotherapy in metastatic pancreatic cancer?

OncoImmunology 5, e1112942 (2016).

33. Watanabe, I. et al. Advanced pancreatic ductal cancer: fibrotic focus and beta-catenin

expression correlate with outcome. Pancreas 26, 326–333 (2003).

34. Kleeff, J. et al. Pancreatic cancer microenvironment. Int J Cancer 121, 699–705 (2007).

35. Hingorani, S. R. et al. HALO 202: Randomized phase II study of PEGPH20 plus nab-

paclitaxel/gemcitabine versus nab-paclitaxel/gemcitabine in patients with untreated,

metastatic pancreatic ductal adenocarcinoma. J Clin Oncol 36, 359–366 (2018).

36. Erkan, M., Kurtoglu, M. & Kleeff, J. The role of hypoxia in pancreatic cancer: a potential

therapeutic target? Expert Rev of Gastroenterol & Hepatol 10, 301–316 (2016).

37. Chang, Q., Jurisica, I., Do, T. & Hedley, D. W. Hypoxia predicts aggressive growth and

spontaneous metastasis formation from orthotopically grown primary xenografts of human

pancreatic cancer. Cancer Res 71, 3110–3120 (2011).

38. Van Cutsem, E. et al. MAESTRO: A randomized, double-blind phase III study of

evofosfamide (Evo) in combination with gemcitabine (Gem) in previously untreated

patients (pts) with metastatic or locally advanced unresectable pancreatic ductal

adenocarcinoma (PDAC). J Clin Oncol 34, 4007–4007 (2016).

39. Dhani, N. C. et al. Analysis of the intra- and intertumoral heterogeneity of hypoxia in

pancreatic cancer patients receiving the nitroimidazole tracer pimonidazole. Br J Cancer

113, 864–871 (2015).

40. Metran-Nascente, C. et al. Measurement of tumor hypoxia in patients with advanced

pancreatic cancer based on 18F-fluoroazomyin arabinoside uptake. J Nucl Med 57, 361–

366 (2016).

41. Almoguera, C. et al. Most human carcinomas of the exocrine pancreas contain mutant c-K-

ras genes. Cell 53, 549–554 (1988).

42. Wilentz, R. E. et al. Inactivation of the p16 (INK4A) tumor-suppressor gene in pancreatic

duct lesions: loss of intranuclear expression. Cancer Res 58, 4740–4744 (1998).

Page 74: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

64

43. Wilentz, R. E. et al. Loss of expression of Dpc4 in pancreatic intraepithelial neoplasia:

evidence that DPC4 inactivation occurs late in neoplastic progression. Cancer Res 60,

2002–2006 (2000).

44. DiGiuseppe, J. A. et al. Overexpression of p53 protein in adenocarcinoma of the pancreas.

Am J Clin Pathol 101, 684–688 (1994).

45. Hruban, R. H., Goggins, M., Parsons, J. & Kern, S. E. Progression model for pancreatic

cancer. Clin Cancer Res 6, 2969–2972 (2000).

46. Notta, F. et al. A renewed model of pancreatic cancer evolution based on genomic

rearrangement patterns. Nature 538, 378–382 (2016).

47. DeNicola, G. M. et al. Oncogene-induced Nrf2 transcription promotes ROS detoxification

and tumorigenesis. Nature 475, 106–109 (2011).

48. Martinez-Useros, J., Li, W., Cabeza-Morales, M. & Garcia-Foncillas, J. Oxidative stress: a

new target for pancreatic cancer prognosis and treatment. J Clin Med 6, 29 (2017).

49. Schieber, M. & Chandel, N. S. ROS function in redox signaling and oxidative stress. Curr

Biol 24, R453–R462 (2014).

50. Sosa, V. et al. Oxidative stress and cancer: an overview. Ageing Res Rev 12, 376–390

(2013).

51. Trachootham, D. et al. Selective killing of oncogenically transformed cells through a ROS-

mediated mechanism by β-phenylethyl isothiocyanate. Cancer Cell 10, 241–252 (2006).

52. Schumacker, P. T. Reactive oxygen species in cancer: a dance with the devil. Cancer Cell

27, 156–157 (2015).

53. Pani, G., Galeotti, T. & Chiarugi, P. Metastasis: cancer cell’s escape from oxidative stress.

Cancer Metastasis Rev 29, 351–378 (2010).

54. Gorrini, C., Harris, I. S. & Mak, T. W. Modulation of oxidative stress as an anticancer

strategy. Nat Rev Drug Discov 12, 931–947 (2013).

55. Vaquero, E. C., Edderkaoui, M., Pandol, S. J., Gukovsky, I. & Gukovskaya, A. S. Reactive

oxygen species produced by NAD(P)H oxidase inhibit apoptosis in pancreatic cancer cells.

J Biol Chem 279, 34643–34654 (2004).

56. Li, X. et al. Targeting mitochondrial reactive oxygen species as novel therapy for

inflammatory diseases and cancers. J Hematol Oncol 6, 19 (2013).

57. Bell, E. L. et al. The Q o site of the mitochondrial complex III is required for the

transduction of hypoxic signaling via reactive oxygen species production. J Cell Biol 177,

1029–1036 (2007).

Page 75: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

65

58. Schrader, M. & Fahimi, H. D. Peroxisomes and oxidative stress. Biochim Biophys Acta

(BBA) - Mol Cell Res 1763, 1755–1766 (2006).

59. Malhotra, J. D. & Kaufman, R. J. Endoplasmic reticulum stress and oxidative stress: a

vicious cycle or a double-edged sword? Antioxid Redox Signal 9, 2277–2293 (2007).

60. Guéraud, F. et al. Chemistry and biochemistry of lipid peroxidation products. Free Radic

Res 44, 1098–1124 (2010).

61. Zarkovic, N. 4-hydroxynonenal as a bioactive marker of pathophysiological processes. Mol

Aspects Med 24, 281–291 (2003).

62. Koritzinsky, M. & Wouters, B. G. The roles of reactive oxygen species and autophagy in

mediating the tolerance of tumor cells to cycling hypoxia. Semin Radiat Oncol 23, 252–261

(2013).

63. Diaz, B. & Yuen, A. The impact of hypoxia in pancreatic cancer invasion and metastasis.

Hypoxia 16, 91-106 (2014).

64. Galadari, S., Rahman, A., Pallichankandy, S. & Thayyullathil, F. Reactive oxygen species

and cancer paradox: to promote or to suppress? Free Radic Biol Med 104, 144–164 (2017).

65. Hengartner, M. O. The biochemistry of apoptosis. Nature 407, 770–776 (2000).

66. Bernardi, P. & Di Lisa, F. The mitochondrial permeability transition pore: molecular nature

and role as a target in cardioprotection. J Mol Cell Cardiol 78, 100–106 (2015).

67. Loreto, C. et al. The role of intrinsic pathway in apoptosis activation and progression in

Peyronie’s Disease. Biomed Res Int 2014, 616149 (2014).

68. Pastorino, J. G., Chen, S. T., Tafani, M., Snyder, J. W. & Farber, J. L. The overexpression

of Bax produces cell death upon induction of the mitochondrial permeability transition. J

Biol Chem 273, 7770–7775 (1998).

69. Dejean, L. M. et al. Oligomeric Bax is a component of the putative cytochrome c release

channel MAC, mitochondrial apoptosis-induced channel. Mol Biol Cell 16, 2424–2432

(2005).

70. Madesh, M. & Hajnóczky, G. VDAC-dependent permeabilization of the outer

mitochondrial membrane by superoxide induces rapid and massive cytochrome c release. J

Cell Biol 155, 1003–1016 (2001).

71. Hedley, D. W., McCulloch, E. A., Minden, M. D., Chow, S. & Curtis, J. Antileukemic

action of buthionine sulfoximine: evidence for an intrinsic death mechanism based on

oxidative stress. Leukemia 12, 1545–1552 (1998).

72. Dixon, S. J. et al. Ferroptosis: an iron-dependent form of nonapoptotic cell death. Cell 149,

1060–1072 (2012).

Page 76: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

66

73. Gout, P. W., Buckley, A. R., Simms, C. R. & Bruchovsky, N. Sulfasalazine, a potent

suppressor of lymphoma growth by inhibition of the x(c)- cystine transporter: a new action

for an old drug. Leukemia 15, 1633–1640 (2001).

74. Dixon, S. J. et al. Pharmacological inhibition of cystine–glutamate exchange induces

endoplasmic reticulum stress and ferroptosis. eLife 3, e02523 (2014).

75. Shaw, A. T. et al. Selective killing of K-ras mutant cancer cells by small molecule inducers

of oxidative stress. Proc Natl Acad Sci U S A 108, 8773–8778 (2011).

76. Terada, Y. et al. Intestinal p-glycoprotein expression is multimodally regulated by intestinal

ischemia-reperfusion. J Pharm Pharm Sci 17, 266–276 (2014).

77. Arora, S. et al. An undesired effect of chemotherapy: gemcitabine promotes pancreatic

cancer cell invasiveness through reactive oxygen species-dependent, nuclear factor κB- and

hypoxia-inducible factor 1α-mediated up-regulation of CXCR4. J Biol Chem 288, 21197–

21207 (2013).

78. Afzal, S. et al. Oxidative damage to guanine nucleosides following combination

chemotherapy with 5-fluorouracil and oxaliplatin. Cancer Chemother Pharmacol 69, 301–

307 (2012).

79. Alexandre, J. et al. Accumulation of hydrogen peroxide is an early and crucial step for

paclitaxel-induced cancer cell death both in vitro and in vivo. Int J Cancer 119, 41–48

(2006).

80. Kotamraju, S., Chitambar, C. R., Kalivendi, S. V., Joseph, J. & Kalyanaraman, B.

Transferrin receptor-dependent iron uptake is responsible for doxorubicin-mediated

apoptosis in endothelial cells: role of oxidant-induced iron signaling in apoptosis. J Biol

Chem 277, 17179–17187 (2002).

81. Zou, Z., Chang, H., Li, H. & Wang, S. Induction of reactive oxygen species: an emerging

approach for cancer therapy. Apoptosis 22, 1321–1335 (2017).

82. Bairati, I. et al. Randomized trial of antioxidant vitamins to prevent acute adverse effects of

radiation therapy in head and neck cancer patients. J Clin Oncol 23, 5805–5813 (2005).

83. Wang, H. et al. Auranofin radiosensitizes tumor cells through targeting thioredoxin

reductase and resulting overproduction of reactive oxygen species. Oncotarget 8, 35728-

35742 (2017).

84. Diehn, M. et al. Association of reactive oxygen species levels and radioresistance in cancer

stem cells. Nature 458, 780–783 (2009).

85. Kansanen, E., Kuosmanen, S. M., Leinonen, H. & Levonen, A. L. The Keap1-Nrf2

pathway: mechanisms of activation and dysregulation in cancer. Redox Biol 1, 45–49

(2013).

Page 77: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

67

86. Itoh, K., Mimura, J. & Yamamoto, M. Discovery of the negative regulator of Nrf2, Keap1:

a historical overview. Antioxid Redox Signal 13, 1665–1678 (2010).

87. Huang, Y., Li, W. & Kong, A. N. T. Anti-oxidative stress regulator NF-E2-related factor 2

mediates the adaptive induction of antioxidant and detoxifying enzymes by lipid

peroxidation metabolite 4-hydroxynonenal. Cell Biosci 2, 40 (2012).

88. Łuczaj, W., Gęgotek, A. & Skrzydlewska, E. Antioxidants and HNE in redox homeostasis.

Free Radic Biol Med 111, 87–101 (2017).

89. Li, C. Q. et al. Nitric oxide activation of Keap1/Nrf2 signaling in human colon carcinoma

cells. Proc Natl Acad Sci U S A 106, 14547–14551 (2009).

90. Isohookana, J., Haapasaari, K. M., Soini, Y. & Karihtala, P. Keap1 expression has

independent prognostic value in pancreatic adenocarcinomas. Diagn Pathol 10, 28 (2015).

91. Sporn, M. B. & Liby, K. T. NRF2 and cancer: the good, the bad and the importance of

context. Nat Rev Cancer 12, 564–571 (2012).

92. Itoh, K. et al. An Nrf2/small Maf heterodimer mediates the induction of phase II

detoxifying enzyme genes through antioxidant response elements. Biochem Biophys Res

Commun 236, 313–322 (1997).

93. Agyeman, A. S. et al. Transcriptomic and proteomic profiling of KEAP1 disrupted and

sulforaphane treated human breast epithelial cells reveals common expression profiles.

Breast Cancer Res Treat 132, 175–187 (2012).

94. Hirotsu, Y. et al. Nrf2–MafG heterodimers contribute globally to antioxidant and metabolic

networks. Nucleic Acids Res 40, 10228–10239 (2012).

95. Yates, M. S. et al. Genetic versus chemoprotective activation of Nrf2 signaling:

overlapping yet distinct gene expression profiles between Keap1 knockout and triterpenoid-

treated mice. Carcinogenesis 30, 1024–1031 (2009).

96. Gutteridge, J. M. C. Iron promoters of the Fenton reaction and lipid peroxidation can be

released from haemoglobin by peroxides. FEBS Letters 201, 291–295 (1986).

97. Taguchi, K., Motohashi, H. & Yamamoto, M. Molecular mechanisms of the Keap1–Nrf2

pathway in stress response and cancer evolution. Genes Cells 16, 123–140 (2011).

98. McGrath-Morrow, S. et al. Nrf2 increases survival and attenuates alveolar growth

inhibition in neonatal mice exposed to hyperoxia. Am J Physiol Lung Cell Mol Physiol 296,

L565–L573 (2009).

99. Chorley, B. N. et al. Identification of novel NRF2-regulated genes by ChIP-Seq: influence

on retinoid X receptor alpha. Nucleic Acids Res 40, 7416–7429 (2012).

100. Warburg, O. On the Origin of Cancer Cells. Science 123, 309–314 (1956).

Page 78: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

68

101. Cairns, R. A., Harris, I. S. & Mak, T. W. Regulation of cancer cell metabolism. Nat Rev

Cancer 11, 85–95 (2011).

102. Gaglio, D. et al. Oncogenic K-Ras decouples glucose and glutamine metabolism to support

cancer cell growth. Mol Syst Biol 7, 523 (2011).

103. Mitsuishi, Y. et al. Nrf2 redirects glucose and glutamine into anabolic pathways in

metabolic reprogramming. Cancer Cell 22, 66–79 (2012).

104. DeNicola, G. M. et al. NRF2 regulates serine biosynthesis in non-small cell lung cancer.

Nat Genet 47, 1475–1481 (2015).

105. Patra, K. C. & Hay, N. The pentose phosphate pathway and cancer. Trends Biochem Sci 39,

347–354 (2014).

106. Hu, T. et al. Variant G6PD levels promote tumor cell proliferation or apoptosis via the

STAT3/5 pathway in the human melanoma xenograft mouse model. BMC Cancer 13, 251

(2013).

107. Ying, H. et al. Oncogenic Kras maintains pancreatic tumors through regulation of anabolic

glucose metabolism. Cell 149, 656–670 (2012).

108. Son, J. et al. Glutamine supports pancreatic cancer growth through a KRAS-regulated

metabolic pathway. Nature 496, 101–105 (2013).

109. Samanta, D. & Semenza, G. L. Serine synthesis helps hypoxic cancer stem cells regulate

redox. Cancer Res 76, 6458–6462 (2016).

110. Jia, X. et al. Increased expression of PHGDH and prognostic significance in colorectal

cancer. Transl Oncol 9, 191–196 (2016).

111. Possemato, R. et al. Functional genomics reveals serine synthesis is essential in PHGDH-

amplified breast cancer. Nature 476, 346–350 (2011).

112. Song, Z., Feng, C., Lu, Y., Lin, Y. & Dong, C. PHGDH is an independent prognosis marker

and contributes cell proliferation, migration and invasion in human pancreatic cancer. Gene

642, 43–50 (2018).

113. Ye, J. et al. Serine catabolism regulates mitochondrial redox control during hypoxia.

Cancer Discov 4, 1406–1417 (2014).

114. Wise, D. R. & Thompson, C. B. Glutamine addiction: a new therapeutic target in cancer.

Trends Biochem Sci 35, 427–433 (2010).

115. DeBerardinis, R. J. et al. Beyond aerobic glycolysis: transformed cells can engage in

glutamine metabolism that exceeds the requirement for protein and nucleotide synthesis.

Proc Natl Acad Sci U S A 104, 19345–19350 (2007).

Page 79: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

69

116. DeBerardinis, R. J. & Cheng, T. Q’s next: the diverse functions of glutamine in

metabolism, cell biology and cancer. Oncogene 29, 313–324 (2010).

117. Cohen, R. et al. Targeting cancer cell metabolism in pancreatic adenocarcinoma.

Oncotarget 6, 16832-16847 (2015).

118. Shi, X., Zhang, Y., Zheng, J. & Pan, J. Reactive oxygen species in cancer stem cells.

Antioxid Redox Signal 16, 1215–1228 (2012).

119. Karisch, R. et al. Global proteomic assessment of the classical protein-tyrosine

phosphatome and “redoxome”. Cell 146, 826–840 (2011).

120. Sato, A. et al. Pivotal role for ROS activation of p38 MAPK in the control of differentiation

and tumor-initiating capacity of glioma-initiating cells. Stem Cell Res 12, 119–131 (2014).

121. Kwon, J. et al. Reversible oxidation and inactivation of the tumor suppressor PTEN in cells

stimulated with peptide growth factors. Proc Natl Acad Sci U S A 101, 16419–16424

(2004).

122. Polytarchou, C., Hatziapostolou, M. & Papadimitriou, E. Hydrogen peroxide stimulates

proliferation and migration of human prostate cancer cells through activation of activator

protein-1 and up-regulation of the heparin affin regulatory peptide gene. J Biol Chem 280,

40428–40435 (2005).

123. Irani, K. et al. Mitogenic signaling mediated by oxidants in Ras-transformed fibroblasts.

Science 275, 1649–1652 (1997).

124. Binker, M. G., Binker-Cosen, A. A., Richards, D., Oliver, B. & Cosen-Binker, L. I. EGF

promotes invasion by PANC-1 cells through Rac1/ROS-dependent secretion and activation

of MMP-2. Biochem Biophys Res Commun 379, 445–450 (2009).

125. Chiu, W.-T. et al. Contribution of reactive oxygen species to migration/invasion of human

glioblastoma cells U87 via ERK-dependent COX-2/PGE(2) activation. Neurobiol Dis 37,

118–129 (2010).

126. Piskounova, E. et al. Oxidative stress inhibits distant metastasis by human melanoma cells.

Nature 527, 186–191 (2015).

127. Harris, I. S. et al. Glutathione and thioredoxin antioxidant pathways synergize to drive

cancer initiation and progression. Cancer Cell 27, 211–222 (2015).

128. Raj, L. et al. Selective killing of cancer cells by a small molecule targeting the stress

response to ROS. Nature 475, 231–234 (2011).

129. Hayes, J. D. & Dinkova-Kostova, A. T. The Nrf2 regulatory network provides an interface

between redox and intermediary metabolism. Trends Biochem Sci 39, 199–218 (2014).

Page 80: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

70

130. Milkovic, L., Zarkovic, N. & Saso, L. Controversy about pharmacological modulation of

Nrf2 for cancer therapy. Redox Biol 12, 727–732 (2017).

131. Ramos-Gomez, M. et al. Sensitivity to carcinogenesis is increased and chemoprotective

efficacy of enzyme inducers is lost in nrf2 transcription factor-deficient mice. Proc Natl

Acad Sci U S A 98, 3410–3415 (2001).

132. Kensler, T. W. et al. Effects of glucosinolate-rich broccoli sprouts on urinary levels of

aflatoxin-DNA adducts and phenanthrene tetraols in a randomized clinical trial in He Zuo

Township, Qidong, People’s Republic of China. Cancer Epidemiol Biomarkers Prev 14,

2605–2613 (2005).

133. Pais, R. & Dumitraşcu, D. L. Do antioxidants prevent colorectal cancer? A meta-analysis.

Rom J Intern Med 51, 152–163 (2013).

134. Goodman, M., Bostick, R. M., Kucuk, O. & Jones, D. P. Clinical trials of antioxidants as

cancer prevention agents: past, present, and future. Free Radic Biol Med 51, 1068–1084

(2011).

135. Sayin, V. I. et al. Antioxidants accelerate lung cancer progression in mice. Sci Transl Med

6, 221ra15 (2014).

136. Klein, E. A. et al. Vitamin E and the risk of prostate cancer: updated results of The

Selenium and Vitamin E Cancer Prevention Trial (SELECT). JAMA 306, 1549–1556

(2011).

137. Yoo, N. J., Kim, H. R., Kim, Y. R., An, C. H. & Lee, S. H. Somatic mutations of the

KEAP1 gene in common solid cancers. Histopathology 60, 943–952 (2012).

138. Konstantinopoulos, P. A. et al. Keap1 mutations and Nrf2 pathway activation in epithelial

ovarian cancer. Cancer Res 71, 5081–5089 (2011).

139. Meister, A. Selective modification of glutathione metabolism. Science 220, 472–477

(1983).

140. Dalton, T. P., Chen, Y., Schneider, S. N., Nebert, D. W. & Shertzer, H. G. Genetically

altered mice to evaluate glutathione homeostasis in health and disease. Free Radic Biol Med

37, 1511–1526 (2004).

141. Seelig, G. F., Simondsen, R. P. & Meister, A. Reversible dissociation of gamma-

glutamylcysteine synthetase into two subunits. J Biol Chem 259, 9345–9347 (1984).

142. Huang, C. S., Anderson, M. E. & Meister, A. Amino acid sequence and function of the light

subunit of rat kidney gamma-glutamylcysteine synthetase. J Biol Chem 268, 20578–20583

(1993).

143. Lauterburg, B. H., Adams, J. D. & Mitchell, J. R. Hepatic glutathione homeostasis in the

rat: efflux accounts for glutathione turnover. Hepatology 4, 586–590 (1984).

Page 81: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

71

144. Grant, C. M., MacIver, F. H. & Dawes, I. W. Glutathione synthetase is dispensable for

growth under both normal and oxidative stress conditions in the yeast Saccharomyces

cerevisiae due to an accumulation of the dipeptide gamma-glutamylcysteine. Mol Biol Cell

8, 1699–1707 (1997).

145. Lewerenz, J. et al. The cystine/glutamate antiporter system xc− in health and disease: from

molecular mechanisms to novel therapeutic opportunities. Antioxid Redox Signal 18, 522–

555 (2013).

146. Higgins, L. G. et al. Transcription factor Nrf2 mediates an adaptive response to

sulforaphane that protects fibroblasts in vitro against the cytotoxic effects of electrophiles,

peroxides and redox-cycling agents. Toxicol Appl Pharmacol 237, 267–280 (2009).

147. Lewerenz, J. et al. Induction of Nrf2 and xCT are involved in the action of the

neuroprotective antibiotic ceftriaxone in vitro. J Neurochem 111, 332–343 (2009).

148. Lu, S. C. Glutathione synthesis. Biochim Biophys Acta (BBA) - Gen Subj 1830, 3143–3153

(2013).

149. Armstrong, J. S. et al. Role of glutathione depletion and reactive oxygen species generation

in apoptotic signaling in a human B lymphoma cell line. Cell Death Differ 9, 252–263

(2002).

150. Garcia‐Ruiz, C. & Fernandez‐Checa, J. C. Mitochondrial glutathione: Hepatocellular

survival–death switch. J Gastroenterol Hepatol 21, S3–S6 (2006).

151. Lu, S. C. Regulation of glutathione synthesis. Mol Aspects Med 30, 42–59 (2009).

152. Paek, J. et al. Mitochondrial SKN-1/Nrf mediates a conserved starvation response. Cell

Metab 16, 526–537 (2012).

153. Harvey, C. J. et al. Nrf2-regulated glutathione recycling independent of biosynthesis is

critical for cell survival during oxidative stress. Free Radic Biol Med 46, 443–453 (2009).

154. Forman, H. J., Zhang, H. & Rinna, A. Glutathione: overview of its protective roles,

measurement, and biosynthesis. Mol Aspects Med 30, 1–12 (2009).

155. Giustarini, D., Dalle-Donne, I., Milzani, A., Fanti, P. & Rossi, R. Analysis of GSH and

GSSG after derivatization with N-ethylmaleimide. Nat Protoc 8, 1660–1669 (2013).

156. Dalle-Donne, I., Rossi, R., Colombo, G., Giustarini, D. & Milzani, A. Protein S-

glutathionylation: a regulatory device from bacteria to humans. Trends Biochem Sci 34, 85–

96 (2009).

157. Zhang, H. & Forman, H. J. Glutathione synthesis and its role in redox signaling. Semin Cell

Dev Biol 23, 722–728 (2012).

Page 82: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

72

158. Townsend, D. M. et al. Novel role for glutathione S-transferase π: regulator of protein S-

glutathionylation following oxidative and nitrosative stress. J Biol Chem 284, 436–445

(2009).

159. Klomsiri, C., Karplus, P. A. & Poole, L. B. Cysteine-based redox switches in enzymes.

Antioxid Redox Signal 14, 1065–1077 (2011).

160. Shelton, M. D., Chock, P. B. & Mieyal, J. J. Glutaredoxin: role in reversible protein s-

glutathionylation and regulation of redox signal transduction and protein translocation.

Antioxid Redox Signal 7, 348–366 (2005).

161. Fratelli, M. et al. Identification of proteins undergoing glutathionylation in oxidatively

stressed hepatocytes and hepatoma cells. Proteomics 3, 1154–1161 (2003).

162. Reynaert, N. L. et al. Dynamic redox control of NF-κB through glutaredoxin-regulated S-

glutathionylation of inhibitory κB kinase β. Proc Natl Acad Sci U S A 103, 13086–13091

(2006).

163. Mohr, S., Hallak, H., Boitte, A. de, Lapetina, E. G. & Brüne, B. Nitric oxide-induced S-

glutathionylation and inactivation of glyceraldehyde-3-phosphate dehydrogenase. J Biol

Chem 274, 9427–9430 (1999).

164. Velu, C. S., Niture, S. K., Doneanu, C. E., Pattabiraman, N. & Srivenugopal, K. S. Human

p53 is inhibited by glutathionylation of cysteines present in the proximal DNA-binding

domain during oxidative stress. Biochemistry 46, 7765–7780 (2007).

165. Hurd, T. R. et al. Glutathionylation of mitochondrial proteins. Antioxid Redox Signal 7,

999–1010 (2005).

166. Allocati, N., Masulli, M., Ilio, C. D. & Federici, L. Glutathione transferases: substrates,

inihibitors and pro-drugs in cancer and neurodegenerative diseases. Oncogenesis 7, 8

(2018).

167. Pasello, M. et al. Overcoming glutathione S-transferase p1–related cisplatin resistance in

osteosarcoma. Cancer Res 68, 6661–6668 (2008).

168. Poot, M., Teubert, H., Rabinovitch, P. S. & Kavanagh, T. J. De novo synthesis of

glutathione is required for both entry into and progression through the cell cycle. J Cell

Physiol 163, 555–560 (2005).

169. Shaw, J. P. & Chou, I. N. Elevation of intracellular glutathione content associated with

mitogenic stimulation of quiescent fibroblasts. Journal of Cellular Physiology 129, 193–

198 (2005).

170. Huang, Z. Z. et al. Mechanism and significance of increased glutathione level in human

hepatocellular carcinoma and liver regeneration. FASEB J 15, 19–21 (2001).

Page 83: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

73

171. Huang, Z. Z. et al. Changes in glutathione homeostasis during liver regeneration in the rat.

Hepatology 27, 147–153 (1998).

172. Carretero, J. et al. Growth-associated changes in glutathione content correlate with liver

metastatic activity of B16 melanoma cells. Clin Exp Metastasis 17, 567–574 (1999).

173. Schnelldorfer, T. et al. Glutathione depletion causes cell growth inhibition and enhanced

apoptosis in pancreatic cancer cells. Cancer 89, 1440–1447 (2000).

174. Diaz Vivancos, P., Wolff, T., Markovic, J., Pallardó, F. V. & Foyer, C. H. A nuclear

glutathione cycle within the cell cycle. Biochem J 431, 169–178 (2010).

175. Hentze, H. et al. Glutathione dependence of caspase-8 activation at the death-inducing

signaling complex. J Biol Chem 277, 5588–5595 (2002).

176. Franklin, C. C. et al. Caspase-3-dependent cleavage of the glutamate-l-cysteine ligase

catalytic subunit during apoptotic cell death. Am J Pathol 160, 1887–1894 (2002).

177. Marengo, B. et al. DNA oxidative damage of neoplastic rat liver lesions. Oncol Rep 23,

1241–1246 (2010).

178. Li, M., Zhang, Z., Yuan, J., Zhang, Y. & Jin, X. Altered glutamate cysteine ligase

expression and activity in renal cell carcinoma. Biomed Rep 2, 831–834 (2014).

179. Habib, E., Linher-Melville, K., Lin, H. X. & Singh, G. Expression of xCT and activity of

system xc− are regulated by NRF2 in human breast cancer cells in response to oxidative

stress. Redox Biol 5, 33–42 (2015).

180. Roomi, M. W. et al. Preneoplastic liver cell foci expansion induced by thioacetamide

toxicity in drug-primed mice. Exp Mol Pathol 81, 8–14 (2006).

181. Obrador, E. et al. γ-Glutamyl transpeptidase overexpression increases metastatic growth of

B16 melanoma cells in the mouse liver. Hepatology 35, 74–81 (2002).

182. Yao, D. et al. Abnormal expression of hepatoma specific γ-glutamyl transferase and

alteration of γ-glutamyl transferase gene methylation status in patients with hepatocellular

carcinoma. Cancer 88, 761–769 (2000).

183. Mulcahy, R. T., Bailey, H. H. & Gipp, J. J. Up-regulation of gamma-glutamylcysteine

synthetase activity in melphalan-resistant human multiple myeloma cells expressing

increased glutathione levels. Cancer Chemother Pharmacol 34, 67–71 (1994).

184. Lewandowicz, G. M. et al. Cellular glutathione content, in vitro chemoresponse, and the

effect of BSO modulation in samples derived from patients with advanced ovarian cancer.

Gynecol Oncol 85, 298–304 (2002).

Page 84: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

74

185. Hidalgo, M. et al. A Phase I and pharmacological study of the glutamine antagonist acivicin

with the amino acid solution aminosyn in patients with advanced solid malignancies. Clin

Cancer Res 4, 2763–2770 (1998).

186. Wickham, S. et al. Inhibition of human gamma-glutamyl transpeptidase: development of

more potent, physiologically relevant, uncompetitive inhibitors. Biochem J 450, 547-557

(2013).

187. Liu, D. S. et al. Inhibiting the system xC−/glutathione axis selectively targets cancers with

mutant-p53 accumulation. Nat Commun 8, 14844 (2017).

188. Sleire, L. et al. Drug repurposing: sulfasalazine sensitizes gliomas to gamma knife

radiosurgery by blocking cystine uptake through system Xc−, leading to glutathione

depletion. Oncogene 34, 5951–5959 (2015).

189. McGirt, M. J. et al. Gliadel (BCNU) wafer plus concomitant temozolomide therapy after

primary resection of glioblastoma multiforme. J Neurosurg 110, 583–588 (2009).

190. Damaj, G. et al. Carmustine replacement in intensive chemotherapy preceding reinjection

of autologous HSCs in Hodgkin and non-Hodgkin lymphoma: a review. Bone Marrow

Transplant 52, 941–949 (2017).

191. Karplus, P. A., Krauth-Siegel, R. L., Schirmer, R. H. & Schulz, G. E. Inhibition of human

glutathione reductase by the nitrosourea drugs 1,3-bis(2-chloroethyl)-1-nitrosourea and 1-

(2-chloroethyl)-3-(2-hydroxyethyl)-1-nitrosourea. A crystallographic analysis. Eur J

Biochem 171, 193–198 (1988).

192. Alessandrino, E. P. et al. Pulmonary toxicity following carmustine-based preparative

regimens and autologous peripheral blood progenitor cell transplantation in hematological

malignancies. Bone Marrow Transplant 25, 309–313 (2000).

193. Weiss, R. B. & Issell, B. F. The nitrosoureas: carmustine (BCNU) and lomustine (CCNU).

Cancer Treat Rev 9, 313–330 (1982).

194. Xie, J., Potter, A., Xie, W., Lynch, C. & Seefeldt, T. Evaluation of a dithiocarbamate

derivative as a model of thiol oxidative stress in H9c2 rat cardiomyocytes. Free Radic Biol

Med 70, 214–222 (2014).

195. Sadhu, S. S., Callegari, E., Zhao, Y., Guan, X. & Seefeldt, T. Evaluation of a

dithiocarbamate derivative as an inhibitor of human glutaredoxin-1. J Enzyme Inhib Med

Chem 28, 456–462 (2013).

196. de Souza, L. F. et al. Inhibition of reductase systems by 2-AAPA modulates peroxiredoxin

oxidation and mitochondrial function in A172 glioblastoma cells. Toxicol in Vitro 42, 273–

280 (2017).

Page 85: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

75

197. Tagde, A., Singh, H., Kang, M. H. & Reynolds, C. P. The glutathione synthesis inhibitor

buthionine sulfoximine synergistically enhanced melphalan activity against preclinical

models of multiple myeloma. Blood Cancer J 4, e229 (2014).

198. Villablanca, J. G. et al. A phase I new approaches to neuroblastoma therapy study of

buthionine sulfoximine and melphalan with autologous stem cells for recurrent/refractory

high-risk neuroblastoma. Pediatr Blood & Cancer 63, 1349–1356 (2016).

199. Bailey, H. H. et al. Phase I study of continuous-infusion L-S,R-buthionine sulfoximine with

intravenous melphalan. J Natl Cancer Inst 89, 1789–1796 (1997).

200. Bailey, H. H. L-S,R-buthionine sulfoximine: historical development and clinical issues.

Chem Biol Interact 111–112, 239–254 (1998).

201. Yang, W. S. et al. Regulation of ferroptotic cancer cell death by GPX4. Cell 156, 317–331

(2014).

202. Lee, H. R. et al. Adaptive response to GSH depletion and resistance to L-buthionine-(S,R)-

sulfoximine: involvement of Nrf2 activation. Mol Cell Biochem 318, 23–31 (2008).

203. Lu, J. & Holmgren, A. The thioredoxin antioxidant system. Free Radic Biol Med 66, 75–87

(2014).

204. Raninga, P. V., Di Trapani, G., Vuckovic, S. & Tonissen, K. F. Cross-talk between two

antioxidants, thioredoxin reductase and heme oxygenase-1, and therapeutic implications for

multiple myeloma. Redox Biol 8, 175–185 (2016).

205. Hawkes, H. J. K., Karlenius, T. C. & Tonissen, K. F. Regulation of the human thioredoxin

gene promoter and its key substrates: a study of functional and putative regulatory

elements. Biochim Biophys Acta 1840, 303–314 (2014).

206. Holmgren, A. Antioxidant function of thioredoxin and glutaredoxin systems. Antioxid

Redox Signal 2, 811–820 (2000).

207. Mustacich, D. & Powis, G. Thioredoxin reductase. Biochem J 346, 1–8 (2000).

208. Fritz-Wolf, K., Kehr, S., Stumpf, M., Rahlfs, S. & Becker, K. Crystal structure of the

human thioredoxin reductase–thioredoxin complex. Nat Commun 2, 383 (2011).

209. Hart, T. et al. High-resolution CRISPR screens reveal fitness genes and genotype-specific

cancer liabilities. Cell 163, 1515–1526 (2015).

210. Zhang, J., Li, X., Han, X., Liu, R. & Fang, J. Targeting the thioredoxin system for cancer

therapy. Trends Pharmacol Sci 38, 794–808 (2017).

211. Zhong, L., Arner, E. S. J. & Holmgren, A. Structure and mechanism of mammalian

thioredoxin reductase: The active site is a redox-active selenolthiol/selenenylsulfide formed

Page 86: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

76

from the conserved cysteine-selenocysteine sequence. Proc Natl Acad Sci U S A 97, 5854–

5859 (2000).

212. Arnér, E. S. J. Focus on mammalian thioredoxin reductases--important selenoproteins with

versatile functions. Biochim Biophys Acta 1790, 495–526 (2009).

213. Zhang, B. et al. Selective selenol fluorescent probes: design, synthesis, structural

determinants, and biological applications. J Am Chem Soc 137, 757–769 (2015).

214. Zhong, L. & Holmgren, A. Essential role of selenium in the catalytic activities of

mammalian thioredoxin reductase revealed by characterization of recombinant enzymes

with selenocysteine mutations. J Biol Chem 275, 18121–18128 (2000).

215. Das, K. C. & Das, C. K. Thioredoxin, a singlet oxygen quencher and hydroxyl radical

scavenger: redox independent functions. Biochem Biophys Res Commun 277, 443–447

(2000).

216. Ernster, L. & Forsmark-Andrée, P. Ubiquinol: an endogenous antioxidant in aerobic

organisms. Clin Investig 71, S60-65 (1993).

217. Xia, L. et al. The mammalian cytosolic selenoenzyme thioredoxin reductase reduces

ubiquinone: a novel mechanism for defense against oxidative stress. J Biol Chem 278,

2141–2146 (2003).

218. Davies, M. J. The oxidative environment and protein damage. Biochim Biophys Acta 1703,

93–109 (2005).

219. Avery, S. V. Molecular targets of oxidative stress. Biochem J 434, 201–210 (2011).

220. Storkey, C., Davies, M. J. & Pattison, D. I. Reevaluation of the rate constants for the

reaction of hypochlorous acid (HOCl) with cysteine, methionine, and peptide derivatives

using a new competition kinetic approach. Free Radic Biol Med 73, 60–66 (2014).

221. Wu, C. et al. Thioredoxin 1-mediated post-translational modifications: reduction,

transnitrosylation, denitrosylation, and related proteomics methodologies. Antioxid Redox

Signal 15, 2565–2604 (2011).

222. Lee, S., Kim, S. M. & Lee, R. T. Thioredoxin and thioredoxin target proteins: from

molecular mechanisms to functional significance. Antioxid Redox Signal 18, 1165–1207

(2013).

223. Rhee, S. G., Woo, H. A., Kil, I. S. & Bae, S. H. Peroxiredoxin functions as a peroxidase

and a regulator and sensor of local peroxides. J Biol Chem 287, 4403–4410 (2012).

224. Yang, K. S. et al. Inactivation of human peroxiredoxin I during catalysis as the result of the

oxidation of the catalytic site cysteine to cysteine-sulfinic acid. J Biol Chem 277, 38029–

38036 (2002).

Page 87: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

77

225. Chae, H. Z., Chung, S. J. & Rhee, S. G. Thioredoxin-dependent peroxide reductase from

yeast. J Biol Chem 269, 27670–27678 (1994).

226. Cox, A. G., Peskin, A. V., Paton, L. N., Winterbourn, C. C. & Hampton, M. B. Redox

potential and peroxide reactivity of human peroxiredoxin 3. Biochemistry 48, 6495–6501

(2009).

227. Marinho, H. S., Antunes, F. & Pinto, R. E. Role of glutathione peroxidase and phospholipid

hydroperoxide glutathione peroxidase in the reduction of lysophospholipid hydroperoxides.

Free Radic Biol Med 22, 871–883 (1997).

228. Nicolussi, A., D’Inzeo, S., Capalbo, C., Giannini, G. & Coppa, A. The role of

peroxiredoxins in cancer. Mol Clin Oncol 6, 139–153 (2017).

229. Wang, T. et al. The role of peroxiredoxin II in radiation-resistant MCF-7 breast cancer

cells. Cancer Res 65, 10338–10346 (2005).

230. Chung, Y. M., Yoo, Y. D., Park, J. K., Kim, Y. T. & Kim, H. J. Increased expression of

peroxiredoxin II confers resistance to cisplatin. Anticancer Res 21, 1129–1133 (2001).

231. Cerda, M. B. et al. Silencing peroxiredoxin-2 sensitizes human colorectal cancer cells to

ionizing radiation and oxaliplatin. Cancer Lett 388, 312–319 (2017).

232. Nordlund, P. & Reichard, P. Ribonucleotide reductases. Annu Rev Biochem 75, 681–706

(2006).

233. Laurent, T. C., Moore, E. C. & Reichard, P. Enzymatic synthesis of deoxyribonucleotides. J

Biol Chem 239, 3436–3444 (1964).

234. Holmgren, A. Hydrogen donor system for Escherichia coli ribonucleoside-diphosphate

reductase dependent upon glutathione. Proc Natl Acad Sci U S A 73, 2275–2279 (1976).

235. Avval, F. Z. & Holmgren, A. Molecular mechanisms of thioredoxin and glutaredoxin as

hydrogen donors for mammalian S phase ribonucleotide reductase. J Biol Chem 284, 8233–

8240 (2009).

236. Yoo, M. H. et al. Targeting thioredoxin reductase 1 reduction in cancer cells inhibits self-

sufficient growth and DNA replication. PLoS One 2, e1112 (2007).

237. Kim, H. Y. & Gladyshev, V. N. Different catalytic mechanisms in mammalian

selenocysteine- and cysteine-containing methionine-R-sulfoxide reductases. PLoS Biol 3,

e375 (2005).

238. Ichijo, H. et al. Induction of apoptosis by ASK1, a mammalian MAPKKK that activates

SAPK/JNK and p38 signaling pathways. Science 275, 90–94 (1997).

Page 88: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

78

239. Liu, Y. & Min, W. Thioredoxin promotes ASK1 ubiquitination and degradation to inhibit

ASK1-mediated apoptosis in a redox activity-independent manner. Circ Res 90, 1259–1266

(2002).

240. Meuillet, E. J., Mahadevan, D., Berggren, M., Coon, A. & Powis, G. Thioredoxin-1 binds

to the C2 domain of PTEN inhibiting PTEN’s lipid phosphatase activity and membrane

binding: a mechanism for the functional loss of PTEN’s tumor suppressor activity. Arch

Biochem Biophys 429, 123–133 (2004).

241. De Preter, G. et al. Inhibition of the pentose phosphate pathway by dichloroacetate unravels

a missing link between aerobic glycolysis and cancer cell proliferation. Oncotarget 7,

2910–2920 (2015).

242. Dunbar, E. M. et al. Phase 1 trial of dichloroacetate (DCA) in adults with recurrent

malignant brain tumors. Invest New Drugs 32, 452–464 (2014).

243. Wang, Q. et al. Rational design of selective allosteric inhibitors of PHGDH and serine

synthesis with anti-tumor activity. Cell Chem Biol 24, 55–65 (2017).

244. Murai, S. et al. Inhibition of malic enzyme 1 disrupts cellular metabolism and leads to

vulnerability in cancer cells in glucose-restricted conditions. Oncogenesis 6, e329 (2017).

245. Wen, Y. et al. Discovery of a novel inhibitor of NAD(P)+-dependent malic enzyme (ME2)

by high-throughput screening. Acta Pharmacol Sin 35, 674–684 (2014).

246. Kirkpatrick, D. L. et al. Mechanisms of inhibition of the thioredoxin growth factor system

by antitumor 2-imidazolyl disulfides. Biochem Pharmacol 55, 987–994 (1998).

247. Baker, A. F. et al. The antitumor thioredoxin-1 inhibitor PX-12 (1-methylpropyl 2-

imidazolyl disulfide) decreases thioredoxin-1 and VEGF levels in cancer patient plasma. J

Lab Clin Med 147, 83–90 (2006).

248. Ramanathan, R. K. et al. A randomized phase II study of PX-12, an inhibitor of thioredoxin

in patients with advanced cancer of the pancreas following progression after a gemcitabine-

containing combination. Cancer Chemother Pharmacol 67, 503–509 (2011).

249. Baker, A. et al. A phase IB trial of 24-hour intravenous PX-12, a thioredoxin-1 inhibitor, in

patients with advanced gastrointestinal cancers. Invest New Drugs 31, 631–641 (2013).

250. Li, C., Peng, Y., Mao, B. & Qian, K. Thioredoxin reductase: a novel, independent

prognostic marker in patients with hepatocellular carcinoma. Oncotarget 6, 17792–17804

(2015).

251. Leone, A., Roca, M. S., Ciardiello, C., Costantini, S. & Budillon, A. Oxidative stress gene

expression profile correlates with cancer patient poor prognosis: identification of crucial

pathways might select novel therapeutic approaches. Oxid Med Cell Longev 2017, 1–18

(2017).

Page 89: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

79

252. Yoo, M. H., Xu, X. M., Carlson, B. A., Gladyshev, V. N. & Hatfield, D. L. Thioredoxin

reductase 1 deficiency reverses tumor phenotype and tumorigenicity of lung carcinoma

cells. J Biol Chem 281, 13005–13008 (2006).

253. Eriksson, S., Prigge, J. R., Talago, E. A., Arnér, E. S. J. & Schmidt, E. E. Dietary

methionine can sustain cytosolic redox homeostasis in the mouse liver. Nat Commun 6,

6479 (2015).

254. Stafford, W. C. et al. Irreversible inhibition of cytosolic thioredoxin reductase 1 as a

mechanistic basis for anticancer therapy. Sci Transl Med 10, eaaf7444 (2018).

255. Roder, C. & Thomson, M. J. Auranofin: repurposing an old drug for a golden new age.

Drugs in R&D 15, 13–20 (2015).

256. Saccoccia, F. et al. Thioredoxin reductase and its inhibitors. Curr Protein Pept Sci 15, 621-

646 (2014).

257. Liu, N. et al. Clinically used antirheumatic agent auranofin is a proteasomal deubiquitinase

inhibitor and inhibits tumor growth. Oncotarget 5, (2014).

258. Rigobello, M. P., Scutari, G., Boscolo, R. & Bindoli, A. Induction of mitochondrial

permeability transition by auranofin, a Gold(I)-phosphine derivative. Br J Pharmacol 136,

1162-1168 (2002).

259. Rigobello, M. P., Folda, A., Baldoin, M. C., Scutari, G. & Bindoli, A. Effect of auranofin

on the mitochondrial generation of hydrogen peroxide. Role of thioredoxin reductase. Free

Radic Res 39, 687–695 (2005).

260. Dai, B. et al. KEAP1-dependent synthetic lethality induced by AKT and TXNRD1

inhibitors in lung cancer. Cancer Res 73, 5532–5543 (2013).

261. Rodman, S. N. et al. Enhancement of radiation response in breast cancer stem cells by

inhibition of thioredoxin- and glutathione-dependent metabolism. Radiat Res 186, 385–395

(2016).

262. Fiskus, W. et al. Auranofin induces lethal oxidative and endoplasmic reticulum stress and

exerts potent preclinical activity against chronic lymphocytic leukemia. Cancer Res 74,

2520–2532 (2014).

263. Hanahan, D. & Weinberg, R. A. Hallmarks of cancer: the next generation. Cell 144, 646–

674 (2011).

264. Hornsveld, M. & Dansen, T. B. The hallmarks of cancer from a redox perspective. Antioxid

Redox Signal 25, 300–325 (2016).

265. Tan, S. X. et al. The thioredoxin-thioredoxin reductase system can function in vivo as an

alternative system to reduce oxidized glutathione in Saccharomyces cerevisiae. J Biol Chem

285, 6118–6126 (2010).

Page 90: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

80

266. Du, Y., Zhang, H., Lu, J. & Holmgren, A. Glutathione and glutaredoxin act as a backup of

human thioredoxin reductase 1 to reduce thioredoxin 1 preventing cell death by

aurothioglucose. J Biol Chem 287, 38210–38219 (2012).

267. Chou, T. C. Drug combination studies and their synergy quantification using the Chou-

Talalay method. Cancer Res 70, 440–446 (2010).

268. You, B. R. & Park, W. H. Auranofin induces mesothelioma cell death through oxidative

stress and GSH depletion. Oncol Rep 35, 546–551 (2016).

269. Karlenius, T. C. & Tonissen, K. F. Thioredoxin and cancer: a role for thioredoxin in all

states of tumor oxygenation. Cancers 2, 209–232 (2010).

270. Fan, C. et al. Enhancement of auranofin-induced lung cancer cell apoptosis by

selenocystine, a natural inhibitor of TrxR1 in vitro and in vivo. Cell Death Dis 5, e1191

(2014).

271. Marzano, C. et al. Inhibition of thioredoxin reductase by auranofin induces apoptosis in

cisplatin-resistant human ovarian cancer cells. Free Radic Biol Med 42, 872–881 (2007).

272. Chen, X. et al. Targeting oxidative stress in embryonal rhabdomyosarcoma. Cancer Cell

24, 710–724 (2013).

273. Wang, T. et al. MYCN drives glutaminolysis in neuroblastoma and confers sensitivity to an

ROS augmenting agent. Cell Death Dis 9, 220 (2018).

274. Franken, N. A. P., Rodermond, H. M., Stap, J., Haveman, J. & van Bree, C. Clonogenic

assay of cells in vitro. Nat Protoc 1, 2315–2319 (2006).

275. Chang, Q. et al. Biodistribution of cisplatin revealed by imaging mass cytometry identifies

extensive collagen binding in tumor and normal tissues. Sci Rep 6, 36641 (2016).

276. Jafari, R. et al. The cellular thermal shift assay for evaluating drug target interactions in

cells. Nat Protoc 9, 2100–2122 (2014).

277. Hedley, D. W. & Chow, S. Evaluation of methods for measuring cellular glutathione

content using flow cytometry. Cytometry 15, 349–358 (1994).

278. Capparelli, E. V., Bricker-Ford, R., Rogers, M. J., McKerrow, J. H. & Reed, S. L. Phase I

clinical trial results of auranofin, a novel antiparasitic agent. Antimicrob Agents Chemother

61, e01947-16 (2017).

279. Bailey, H. H. et al. Phase I clinical trial of intravenous L-buthionine sulfoximine and

melphalan: an attempt at modulation of glutathione. J Clin Oncol 12, 194–205 (1994).

280. Topkas, E. et al. Auranofin is a potent suppressor of osteosarcoma metastasis. Oncotarget

7, 831-844 (2016).

Page 91: Targeting Antioxidant Systems to Treat Pancreatic Cancer · immune checkpoint blockade, inhibition of immunosuppressive signalling, stimulation of T cells, and/or conventional cytotoxic

81

281. Blocka, K. Auranofin versus injectable gold. Am J Med 75, 114–122 (1983).

282. Forum on Neuroscience and Nervous System Disorders; Board on Health Sciences Policy;

Institute of Medicine. Improving and Accelerating Therapeutic Development for Nervous

System Disorders: Workshop Summary. Washington (DC): National Academies Press

(US); 2014 Feb 6. 2, Drug Development Challenges. Available from:

https://www.ncbi.nlm.nih.gov/books/NBK195047/.

283. Murphy, M. P. How mitochondria produce reactive oxygen species. Biochem J 417, 1–13

(2009).

284. Yang, W. S. & Stockwell, B. R. Ferroptosis: death by lipid peroxidation. Trends Cell Biol

26, 165–176 (2016).

285. Molavian, H. R. et al. Drug-induced reactive oxygen species (ROS) rely on cell membrane

properties to exert anticancer effects. Sci Rep 6, 27439 (2016).

286. Kean, W. F., Hart, L. & Buchanan, W. Auranofin. Br J Rheumatol 36, 560–572 (1997).

287. Walz, D. T., DiMartino, M. J., Griswold, D. E., Intoccia, A. P. & Flanagan, T. L. Biologic

actions and pharmacokinetic studies of auranofin. Am J Med 75, 90–108 (1983).

288. Drummen, G. P., van Liebergen, L. C., Op den Kamp, J. A. & Post, J. A. C11-

BODIPY581/591, an oxidation-sensitive fluorescent lipid peroxidation probe:

(micro)spectroscopic characterization and validation of methodology. Free Radic Biol Med

33, 473–490 (2002).

289. Rahman, I., Kode, A. & Biswas, S. K. Assay for quantitative determination of glutathione

and glutathione disulfide levels using enzymatic recycling method. Nat Protoc 1, 3159–

3165 (2007).

290. Holmgren, A. Thioredoxin catalyzes the reduction of insulin disulfides by dithiothreitol and

dihydrolipoamide. J Biol Chem 254, 9627–9632 (1979).

291. Arnér, E. S. J. & Holmgren, A. Measurement of thioredoxin and thioredoxin reductase.

Curr Protoc Toxicol 24, 7.4.1-7.4.14 (2005).