Exploring Pancreatic Metabolism and Malignancy [1st ed. 2019] 978-981-32-9392-2, 978-981-32-9393-9

This book comprehensively describes the association between metabolic syndrome and pancreatic cancer progression, and th

409 78 6MB

English Pages XII, 268 [269] Year 2019

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Exploring Pancreatic Metabolism and Malignancy [1st ed. 2019]
 978-981-32-9392-2, 978-981-32-9393-9

Table of contents :
Front Matter ....Pages i-xii
Biology of Pancreas and Possible Diseases (Gowru Srivani, Begum Dariya, Batoul Farran, Afroz Alam, Ganji Purnachandra Nagaraju)....Pages 1-25
Pancreatitis: Clinical Aspects of Inflammatory Phenotypes (Nyshadham S. N. Chaitanya, Aramati BM Reddy)....Pages 27-33
Diabetes and Pancreatic Cancer: A Bidirectional Relationship Perspective (Manoj Kumar Gupta, Vemula Sarojamma, Ramakrishna Vadde)....Pages 35-51
Metabolic Adaptations in Diabetes Mellitus and Cancer (Anil Kumar Pasupulati, Nageswara Rao Dunna, Srikanth Talluri)....Pages 53-69
Role of Mitochondria in Pancreatic Metabolism, Diabetes, and Cancer (Noble Kumar Talari, Ushodaya Mattam, Naresh Babu V. Sepuri)....Pages 71-94
Targeting Mitochondrial Enzymes in Pancreatic Cancer (Gowru Srivani, Begum Dariya, Afroz Alam, Ganji Purnachandra Nagaraju)....Pages 95-110
Diabetes with Pancreatic Ductal Adenocarcinoma (Gowru Srivani, Begum Dariya, Afroz Alam, Ganji Purnachandra Nagaraju)....Pages 111-131
Role of Inflammatory Cytokines in the Initiation and Progression of Pancreatic Cancer (Madanraj Appiya Santharam, Vignesh Dhandapani)....Pages 133-156
Perspectives and Molecular Understanding of Pancreatic Cancer Stem Cells (L. Saikrishna, Prameswari Kasa, Saimila Momin, L. V. K. S. Bhaskar)....Pages 157-172
The Role of Hypoxia Inducible Factor-1α in Pancreatic Cancer and Diabetes Mellitus (Saimila Momin, Ganji Purnachandra Nagaraju)....Pages 173-181
Role of Heat Shock Protein 90 in Diabetes and Pancreatic Cancer Management (Pinninti Santosh Sushma, Saimila Momin, Gowru Srivani)....Pages 183-195
Insulin Resistance Is a Common Core Tethered to Diabetes and Pancreatic Cancer Risk (Henu Kumar Verma, L. V. K. S. Bhaskar)....Pages 197-213
Immunotherapy for Diabetogenic Pancreatitis and Pancreatic Cancer: An Update (Sathish Kumar Mungamuri, Anil Kumar Pasupulati, Vijay Aditya Mavuduru)....Pages 215-236
Exosomes: Mediators and Therapeutic Targets of Diabetes and Pancreatic Cancer ( Deepak KGK, Rama Rao Malla)....Pages 237-251
Methods and Models in Exploring Pancreatic Functions (Rama Rao Malla, Seema Kumari, Krishna Chaitanya Amajala, Deepak KGK, Shailender Gugalavath, Prasuja Rokkam)....Pages 253-268

Citation preview

Ganji Purnachandra Nagaraju  Aramati BM Reddy Editors

Exploring Pancreatic Metabolism and Malignancy

Exploring Pancreatic Metabolism and Malignancy

Ganji Purnachandra Nagaraju Aramati BM Reddy Editors

Exploring Pancreatic Metabolism and Malignancy

Editors Ganji Purnachandra Nagaraju Department of Hematology and Medical Oncology Emory University School of Medicine Atlanta, GA, USA

Aramati BM Reddy Department of Animal Biology University of Hyderabad Hyderabad, Telangana, India

ISBN 978-981-32-9392-2 ISBN 978-981-32-9393-9 https://doi.org/10.1007/978-981-32-9393-9

(eBook)

# Springer Nature Singapore Pte Ltd. 2019 This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. The publisher, the authors, and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, expressed or implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. This Springer imprint is published by the registered company Springer Nature Singapore Pte Ltd. The registered company address is: 152 Beach Road, #21-01/04 Gateway East, Singapore 189721, Singapore

Preface

Notably, roughly 50% of patients with newly diagnosed pancreatic cancer have an existing diabetes diagnosis. Despite several studies evaluating the association between pancreatic adenocarcinoma and diabetes mellitus type 2, the direction of the relationship is not clear. Some studies report that diabetic patients have a two-fold increased risk of developing pancreatic cancer, while others suggest that increased risk of developing cancer exists in newly diagnosed diabetics and progressively decreases with a prolonged course of diabetes mellitus type 2. Several meta-analyses aiming to analyze this relationship have reached various conclusions; in general, they conclude that patients with diabetes mellitus type 2 are likely to develop pancreatic cancer 24–36 months following diagnosis. This relationship, diabetes causing PC development, has been thoroughly evaluated. On the other hand, diabetes development following PC diagnosis is thought to be a result of tumor infiltration, ductal obstruction, and endocrine glandular destruction. PC’s release of mediators affects insulin secretion and action. Therefore, such patients typically demonstrate insulin resistance and hyperinsulinemia. Further studies must be done to elucidate this bidirectional relationship. In this book, we aim to discuss methods and models for exploring pancreatic functions, pancreatic biology, other pancreatic diseases, the role of various organelles, therapeutic options for diabetes mellitus type 2 and pancreatic cancer, and pancreatic functions. The pancreas is a complex organ with multiple functions; it functions as an exocrine and endocrine gland and plays a role in many important pathways. A thorough understanding of pancreas anatomy and pancreatic cancer biology is key for developing new, more effective medications for pancreatic diseases including diabetes mellitus type 2 and pancreatitis, which have been shown to contribute to the development of pancreatic cancer. Their relationship is explored in this book. Mitochondria and exosomes are also important mediators of diabetes and pancreatic cancer. As organelles play a key role in the development of cancer, they may be targeted in future pancreatic cancer therapy. In this book, we will discuss current and innovative therapeutic options for diabetes and pancreatic cancer. Atlanta, GA, USA Hyderabad, Telangana, India

Ganji Purnachandra Nagaraju Aramati BM Reddy

v

Acknowledgment

This book is dedicated to our families, teachers, contributors, and friends.

vii

Contents

1

Biology of Pancreas and Possible Diseases . . . . . . . . . . . . . . . . . . . . Gowru Srivani, Begum Dariya, Batoul Farran, Afroz Alam, and Ganji Purnachandra Nagaraju

1

2

Pancreatitis: Clinical Aspects of Inflammatory Phenotypes . . . . . . . Nyshadham S. N. Chaitanya and Aramati BM Reddy

27

3

Diabetes and Pancreatic Cancer: A Bidirectional Relationship Perspective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Manoj Kumar Gupta, Vemula Sarojamma, and Ramakrishna Vadde

4

Metabolic Adaptations in Diabetes Mellitus and Cancer . . . . . . . . . Anil Kumar Pasupulati, Nageswara Rao Dunna, and Srikanth Talluri

5

Role of Mitochondria in Pancreatic Metabolism, Diabetes, and Cancer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Noble Kumar Talari, Ushodaya Mattam, and Naresh Babu V. Sepuri

35 53

71

6

Targeting Mitochondrial Enzymes in Pancreatic Cancer . . . . . . . . . Gowru Srivani, Begum Dariya, Afroz Alam, and Ganji Purnachandra Nagaraju

95

7

Diabetes with Pancreatic Ductal Adenocarcinoma . . . . . . . . . . . . . . 111 Gowru Srivani, Begum Dariya, Afroz Alam, and Ganji Purnachandra Nagaraju

8

Role of Inflammatory Cytokines in the Initiation and Progression of Pancreatic Cancer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133 Madanraj Appiya Santharam and Vignesh Dhandapani

9

Perspectives and Molecular Understanding of Pancreatic Cancer Stem Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157 L. Saikrishna, Prameswari Kasa, Saimila Momin, and L. V. K. S. Bhaskar

ix

x

Contents

10

The Role of Hypoxia Inducible Factor-1α in Pancreatic Cancer and Diabetes Mellitus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173 Saimila Momin and Ganji Purnachandra Nagaraju

11

Role of Heat Shock Protein 90 in Diabetes and Pancreatic Cancer Management . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183 Pinninti Santosh Sushma, Saimila Momin, and Gowru Srivani

12

Insulin Resistance Is a Common Core Tethered to Diabetes and Pancreatic Cancer Risk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197 Henu Kumar Verma and L. V. K. S. Bhaskar

13

Immunotherapy for Diabetogenic Pancreatitis and Pancreatic Cancer: An Update . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215 Sathish Kumar Mungamuri, Anil Kumar Pasupulati, and Vijay Aditya Mavuduru

14

Exosomes: Mediators and Therapeutic Targets of Diabetes and Pancreatic Cancer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237 Deepak KGK and Rama Rao Malla

15

Methods and Models in Exploring Pancreatic Functions . . . . . . . . . 253 Rama Rao Malla, Seema Kumari, Krishna Chaitanya Amajala, Deepak KGK, Shailender Gugalavath, and Prasuja Rokkam

About the Editors

Ganji Purnachandra Nagaraju is a faculty member at the Department of Hematology and Medical Oncology at Emory University School of Medicine. He obtained his M.Sc. and Ph.D., both in biotechnology from Sri Venkateswara University in Tirupati, Andhra Pradesh, India. He received his D.Sc. from Berhampur University in Berhampur, Odisha, India. His research focuses on translational research in gastrointestinal malignancies. He has published over 70 research papers in highly respected international journals and has presented more than 50 abstracts at various national and international conferences. He is author and editor of several. He is an editorial board member of several internationally recognized academic journals, and is an associate member of the Discovery and Developmental Therapeutics research program at Winship Cancer Institute. He is a member of the Association of Scientists of Indian Origin in America (ASIOA), the Society for Integrative and Comparative Biology (SICB), the Science Advisory Board, the RNA Society and the American Association of Cancer Research (AACR). Aramati BM Reddy is an Assistant Professor at the School of Life Sciences, University of Hyderabad, India. He obtained his M.Sc. in Biotechnology from SV University, Tirupati, India, and Ph.D. from L.V. Prasad Eye Institute, University of Hyderabad, India. He received postdoctoral training at the University of Pennsylvania and University of Texas Medical Branch, Galveston, in epigenetics, inflammation, and metabolic disorders. His research interests focus on understanding cell signaling and gene regulation in the pathophysiology of metabolic malignancies and glaucoma. His lab is xi

xii

About the Editors

also working on understanding the role of several candidate genes and their significance for the outflow pathway of the aqueous humor of the anterior segment of the eye and glaucoma pathology. He has published 30 research papers in peer-reviewed journals and four book chapters. In addition, he has co-edited three special issues for Hindawi publishers in the field of translational and health sciences. He is a life member for several societies and is currently an executive board member of the Indian Society of Cell Biology.

1

Biology of Pancreas and Possible Diseases Gowru Srivani, Begum Dariya, Batoul Farran, Afroz Alam, and Ganji Purnachandra Nagaraju

Abstract

Pancreas is one of the small, flattened organs in the gastrointestinal tract associated with dynamic functions. Most of the food is digested before entering into the other organs and then is converted into energy to regulate the metabolic systems. However, digestion mainly involves the hydrolysis of carbohydrates, lipids, amino acids, proteins, and fatty acids. Releasing enzymes and hormones, which are crucial for regulating the glucose homeostasis and energy production, is the main function of both the endocrine and exocrine pancreas. Disorders of the pancreas can affect the exocrine and endocrine functions, thus resulting in devastating diseases, such as acute and chronic pancreatitis, diabetes, and pancreatic cancer. The main predisposing factors for the development and progression of these disorders are still unclear; therefore, it is important to understand the mechanisms that regulate pancreas homeostasis. In this present chapter, we discuss the biology of the pancreas and its functions and the development of disorders and pathological conditions. We also examine how this accumulating knowledge is guiding the discovery of new diagnostic and therapeutic approaches to cure or prevent disease. Keywords

Pancreas · Exocrine · Endocrine · Pancreatitis · Diabetes and pancreatic cancer

G. Srivani · B. Dariya · A. Alam Department of Bioscience and Biotechnology, Banasthali University, Vanasthali, Rajasthan, India B. Farran · G. P. Nagaraju (*) Department of Hematology and Medical Oncology, Winship Cancer Institute, Emory University, Atlanta, GA, USA e-mail: [email protected] # Springer Nature Singapore Pte Ltd. 2019 G. P. Nagaraju, A. BM Reddy (eds.), Exploring Pancreatic Metabolism and Malignancy, https://doi.org/10.1007/978-981-32-9393-9_1

1

2

G. Srivani et al.

Abbreviations 4EBP1 Akt AMPK ASC Atg1 Atg5 Atg6 ATP CD CDK4I CDKN2A COSMIC DAMPs DNA EGFR ERK FOXO FTIs G1 phase GAD65 GAD67 GTP H HIF-1α HMGB1 HSP70 ICA-512 IDF IL-33 IL-1 IL-10 IL-18 IL-6 IL-8 IL-β JNK MAPK MEK MEN1 mTOR NETS NF-Kb

4E-Binding protein Protein kinase B Adenosine monophosphate-activated protein kinase Apoptosis-associated speck-like protein Autophagy-related 1 Autophagy-related 5 Autophagy-related 6 Adenosine triphosphate Cluster of differentiation Cyclin-dependent kinase 4 inhibitor Cyclin-dependent kinase inhibitor 2A Catalogue of Somatic Mutations in Cancer Damage-associated molecular patterns Deoxyribonucleic acid Epidermal growth factor receptor Extracellular signal-regulated kinases Forkhead box protein Farnesyltransferase inhibitors Gap 1 phase Glutamic acid decarboxylase 65 Glutamic acid decarboxylase 67 Guanosine-5’-triphosphate Histidine Hypoxia-inducible factor-1α High-mobility group box protein 1 Heat-shock protein 70 Islet cell antigen 512 International Diabetes Foundation Interleukin-33 Interleukin-1 Interleukin-10 Interleukin-18 Interleukin-6 Interleukin-8 Interleukin-β c-Jun amino N-terminal kinase Mitogen-activated protein kinase Mitogen-activated protein kinase Multiple endocrine neoplasia Mammalian target of rapamycin Neuroendocrine tumors Nuclear factor kappa B

1

Biology of Pancreas and Possible Diseases

NLRP3 P16 p38 PAMPs PanIN Ph domain PI3K PIP3 PKC PLC PNEC R RAS Rictor RNA ROS S phase SMAD4 SPINK1 TLR TNF TNF-α Tp53 Ulk1 gene VEGF VEGFR VHL

3

Nucleotide-binding domain, leucine-rich repeat-containing family, pyrin domain-containing-3 Protein 16 Protein 38 Pathogen-associated molecular patterns Pancreatic intraepithelial neoplasia Pleckstrin homology Phosphatidylinositol 3-kinase Phosphatidylinositol 3, 4, 5-triphosphate Protein kinase C Phospholipase C Pancreatic necrosis Arginine Reticular activating system Rapamycin-insensitive companion of mTOR Ribonucleic acid Reactive oxygen species Synthesis phase Small worm phenotype and Drosophila mothers against decapentaplegic 4 Serine protease inhibitor Kazal type 1 gene Toll like receptors Tumor necrosis factor Tumor necrosis factor-α Tumor protein 53 Unc-51 like autophagy activating kinase Vascular endothelial growth factor Vascular endothelial growth factor receptor von Hippel–Lindau

1.1

Introduction

1.1.1

Pancreas Location and Anatomy

The pancreas is a glandular and digestive retroperitoneal organ, positioned at the point of the second lumber vertebra, and lies behind the stomach in the abdomen. It is closely related with the duodenum as both organs share a common blood supply. The pancreas structurally consists of the following regions: 1-. Head – A rounded and large portion of the head is placed on the abdomen’s right side and adjacent to the C loop of the duodenum, which receives the pancreatic fluid delivered by the pancreatic duct.

4

G. Srivani et al.

2-. Neck – It is the smallest portion of the pancreas and is directly situated in the superior mesenteric vessels and portal vein. 3-. Body – It is wrapped behind the stomach and is surrounded by the splenic artery and vein. 4-. Tail – The last part of the pancreas and continuation of the body. Both are extended at the level of the thoracic vertebra [1].

1.2

Functions of Pancreas

The pancreas plays a vital role in the production of digestive enzymes as well as glucose homeostasis and is also capable of both endocrine and exocrine functions. The endocrine function of the pancreas is mediated by a cluster of tiny cells distributed on the center known as the islets of Langerhans, which consist of four kinds of cells: the α-cells, β-cells, δ-cells, and γ-cells. These cells secrete insulin, somatostatin, glucagon, and pancreatic polypeptides, respectively. The α-cells and β-cells play a crucial part in maintaining carbohydrate metabolism, lipids, proteins, and fats as well as glucose levels in the blood (Table 1.1). The exocrine function of the pancreas is to secrete pancreatic juices, which are constricted with digestive enzymes such as bicarbonates, proteases, lipases, and amylases that help in the breakdown of food (Table 1.2) [1].

1.3

Diseases of the Pancreas

The devastating nature of pancreatic diseases is mainly due to metabolic imbalances such as congenital diseases and endocrine functional disorders, which can cause diabetes mellitus and pancreatic cancer (PC), one of the deadliest cancers worldwide. They can also be caused by the dysregulation of the exocrine pancreas, resulting in conditions such as pancreatitis, exocrine pancreatic insufficiency, and pancreatic adenocarcinoma.

1.4

Pancreatitis

Pancreatitis is defined as an inflammatory damage of the pancreas caused by excess alcohol consumption, smoking, certain genetic alterations, toxins, bile duct blocked by the gallstones, and sequential activation of pancreatic proenzymes, which are secreted by acinar cells such as proelastase, prophospholipase A2, procolipase, and chymotrypsinogen. These proenzymes are stored in the form of zymogens. There are two forms of pancreatitis: acute and chronic pancreatitis. Acute pancreatitis – In majority of the cases, acute pancreatitis inflammation is driven by several etiological factors (Fig. 1.1) that damage the tissue. Many of these factors are very rare, while others play a significant role in causing acute pancreatitis. Although 40% of the cases are biliary and 30% are alcoholic, the remaining

Insulin (51 amino acid peptide)

Somatostatin (14 amino acid peptide)

Pancreatic polypeptide (36 amino acid peptide)

δ-Cells

γ-Cells

Hormones Glucagon (29 amino acid peptide)

β-Cells

Islet cells α-Cells

PDX1 and PAX4

NKX6.1, NKX2.2, MAFB, PDX1, and PAX6

Transcriptional factors MAFB and PAX6

1

10

70

Composition (%) 20

Table 1.1 Endocrine pancreas – islet of Langerhans cells (Janet L. Fun) Function Glucagon activates gluconeogenesis (glucose synthesis) and glycogenolysis (glycogen storing), actively participates in fatty acid oxidation and ketogenesis (generates ketone bodies) when glucose is not available in the brain (Richard N. Mitchell 2014) Glucose homeostasis and metabolism; it promotes the synthesis of glucose transporter-4 (GLUT-4) in the adipose tissue, cardiac muscle, and skeletal muscle Acts as autocrine function and shows the paracrine function on islet cells producing hormones (suppresses the glucagon and insulin release) in the gastrointestinal tract Somatostatin slows down the nutrient absorption through suppression of gut motility and suppression of exocrine pancreatic function Controls both the endocrine and exocrine functions

Gastrointestinal tract

Brain, pituitary, pancreas, and gastrointestinal tract – mainly gut

Liver, muscle, fat

Targeted tissue Liver

1 Biology of Pancreas and Possible Diseases 5

6

G. Srivani et al.

Table 1.2 Exocrine functions Secretion enzymes (inactive forms) Trypsinogen

Activator Enterokinase

Enzyme Trypsin

Chymotrypsinogen

Trypsin

Chymotrypsin

Proprotease E

Protease E

Procarboxypeptidase A Procarboxypeptidase B

Carboxypeptidase A Carboxypeptidase B

Proelastase 2

Elastase 2

Procolipase

Colipase

Prophospholipase A2

Phospholipids

Ribonuclease Deoxyribonuclease Pancreatic lipase

RNA DNA Triglycerides

Function Breakdown of peptide bonds in aromatic amino acids Catalysis: the hydrolysis of peptide bonds and ester bonds; cleaving peptide bonds at carboxyl side chain of aromatic amino acids Breakdown of proteins, lipids, and carbohydrates Breakdown of carboxy terminal amino acids with aromatic rings Breakdown of carboxy terminal amino acids with basic side chain Breakdown of bonds at carboxyl side of the aliphatic amino acids Forming the active site for pancreatic lipase Catalysis: the hydrolysis of the fatty acyl bond to release fatty acids and lysophospholipids; precursor molecule for eicosanoides Formation of nucleotides Formation of nucleotides Cleaving the dietary fatty acids into fatty acids and glycerol

conditions are associated with other factors. Initially, these factors alter the physiology of the pancreas, resulting in the activation of pathological contents including the autodigestion of enzymes in acinar cells, which are responsible for inflammation [2]. In acute pancreatitis, pancreatic necrosis (PNEC) is a major complication leading to morbidity in a short time span. In all cases of acute pancreatitis, alcohol abuse contributes to 35% of the conditions and is the major risk factor. In fact, chronic alcohol exposure can result in the development of PNEC. Patients suffer from subclinical and multiple clinical disorders during the sequential development of chronic pancreatitis. In fact, chronic alcohol ingestion induces the release of the protein-rich pancreatic fluid, which develops the insoluble protein plugs that cause difficulties in the pancreatic ductules [2]. Hypocalcemia (increased levels of serum calcium) promotes the intraglandular activation action of zymogens, caused by the deposition of calcium (calcium salts) in the pancreatic ducts, and releases fatty acids, leading to hyperparathyroidism [3]. Hereditary pancreatitis occurs in both acute pancreatitis and chronic pancreatitis and is associated with germline mutations in the PRESS1 gene (cationic trypsinogen

1

Biology of Pancreas and Possible Diseases

7

Gall stones

Alcohol, Smoking

Hyperglycemia Hyper calcaemia Hypothermia

Acute Pancreatitis

ERCP

Drugs

Hyperlipidemia

Gene alterations Toxins

Microbiome infections – cocksackie, measles, mycoplasma, mumps, helicobacter pylori, hepatitis and epstein bar virus

Ischaemia

Fig. 1.1 Etiological factors for acute pancreatitis; the main risk factors for the development of acute pancreatitis are gallstones, hyperglycemia, hypercalcemia, toxins, drugs gene alterations, and ischemia. Other contributing factors include microbiome infections – measles, mycoplasma, mumps, helicobacter pylori, hepatitis, and Epstein–Barr virus

gene), namely, point mutations. For instance, a single guanine to adenine (G to A) transition was detected in the 3rd (third) exon of the PRESS1 gene, resulting in an R– H (arginine to histidine) substitution termed R117H. Another mutation results from a single A–T transition identified at amino acid 21, leading to an asparagine to isoleucine substitution termed N211. These mutations inactivate the trypsinogen and trypsin activity. The resulting mutated trypsin abnormally activates the other proenzymes, causing the development of acute pancreatitis [4]. Additionally, mutations in the gene serine protease inhibitor Kazal type 1 (SPINK1), a trypsin inhibitor, can lead to the development of pancreatitis [5]. Obesity has been shown to contribute to acute pancreatitis as well as chronic pancreatitis. In fact, an elevated level of obesity increases the leptin, which is responsible for pro-inflammatory conditions, and decreases the anti-inflammatory agent adiponectin [6]. Furthermore, obesity enhances the number of CD8+ T-cells in macrophages and downregulates T-regulatory cells, further elevating the levels of inflammatory intermediaries such as IL-1β, IL-8, IL-6, and tumor necrosis factor-α (TNF-α) which activate the inflammasome and can lead to inflammation [7]. Moreover, through activating the Akt–mTOR signaling pathway, obesity suppresses the autophagy genes Atg5, Atg6/Beclin1, and Ulk1/Atg1, which increases the severity of pancreatitis [8]. Chronic pancreatitis causes the permanent loss of the pancreas due to irreversible damage, ductal dilation, fibrosis, and damage of acinar cell and islets tissue [4]. The development of acute and chronic pancreatitis is triggered by inflammation driven by the acinar cell damage due to the activation of intrapancreatic trypsinogen, although

8

G. Srivani et al.

Alcohol, smoking Gallstones Protein Kinase C Hyper glycemia Hyper calcaemia

NLR3 Inflammasome Inflammation

Activates NF-κB

Activates the STAT3 Genetic Mutations PRESS1, SPINK1

DAMPS &PAMPS TLR3 /TLR4

ROS Produces IL6,IL10,TNFα

Fibrosis

Fat accumulation Activates oncogens Kras

Acute Pancreatitis Chronic Pancreatitis Diabetes Fig. 1.2 Underlying mechanism of pancreatic diseases. Inflammation plays a central role in initiating the changes in the cellular microenvironment and thus results in the progression of diseases from acute pancreatitis to chronic pancreatitis, diabetes, and pancreatic cancer. Risk factors (alcohol, smoking) initiate the inflammation, and mutation of the genes activates the PRESS1 and SPINK1 genes; inflammatory mediators IL-6 and IL-10 cause the fibrosis and fat accumulation, activate the oncogene KRAS, and promote the pancreatic cancer

acinar cells sense tissue damage through several intracellular molecules released to the extracellular surface, which provoke damage-associated molecular patterns and pathogen-associated molecular patterns (DAMPs and PAMPs). However, a number of DAMPs and PAMPs are triggered by inflammation via adenosine triphosphate (ATP); high-mobility group box protein 1 (HMGB1); ROS (reactive oxygen species); heat-shock protein (HSP70); toll like receptors (TLR)-4; TLR-9; saturated fatty acids; innate immune molecules including IL-6, IL-10, IL-18, IL-β, IL-1, and IL-33; tumor necrosis factor; and protease cathepsin B, shown to play a significant role in experimental acute and chronic pancreatitis. These pathways stimulate NF-κB activation and express the NLRP3 inflammasome, which contains the NLRP3 (nucleotide-binding domain, leucine-rich repeat-containing family, pyrin domaincontaining-3) protein complex and procaspase 1 and ASC. Figure 1.2 (Apoptosisassociated speck-like protein) [9, 10] A long-term standing of chronic pancreatitis is the development of diabetes mellitus. However, diabetes occurs due to the inflammation that destroys the islet cells such as β- and α-cells, undigested fat accumulation, fibrosis in the pancreas,

1

Biology of Pancreas and Possible Diseases

9

deficiency in vitamins (A, D, E, and K), and inflammatory cell infiltration that can damage the pancreatic structure and function [11].

1.5

Diabetes Mellitus

Diabetes mellitus is characterized by endocrine or metabolic disorders including anabolism and catabolism of carbohydrates, fatty acids, and lipid; protein dysregulation of regulatory networks responsible for storage and the mobilization of metabolic functions; glucose intolerance and hyperglycemia leading to improper insulin action; failure of insulin secretion and insulin resistance; or combined deficiency [12]. According to the International Diabetes Foundation (IDF), the number of people between the age group 20 and 79 years suffering from diabetes worldwide will raise from 415 million in 2015 to 642 million in 2040, suggesting that one in five people at least will suffer from diabetes [13]. Based on its mechanisms and clinical symptoms, diabetes mellitus is classified into two types: type 1 diabetes and type 2 diabetes [14]. Furthermore, based on etiology and clinical symptoms, diabetes mellitus can be further classified into four categories: type 1 diabetes, type 2 diabetes, gestational diabetes, and other specific types [15]. Type 1 diabetes is also termed as insulin-dependent, juvenile-onset, and T-cellmediated diabetes. The main causative factors for its development include inflammation of the pancreas, deficiency of insulin through autoimmune reaction leading to the β-cell destruction produced by the islets of exocrine pancreatic cells, and infiltration of leucocytes, i.e., natural killer cells (NK cells), B-cells, and CD8–Tcells [14, 16]. Type 1 diabetes displays mainly four symptoms, namely, polyphagia, glycosuria, polyuria, and polydipsia. Usually, type 1 diabetes occurs in children below 4 years of age which account for the highest incidental reports. Fifty percent of the patients are approximately 20 years old and present with fasting hyperglycemia that leads to ketoacidosis and weight loss and polyphagia among other conditions (American Diabetes Association 2014). Furthermore, about 80% of patients express elevated levels of antibodies of the islet cell antibodies including autoantibodies to insulin/proinsulin, GAD65 and GAD67 (glutamic acid decarboxylase), tyrosine phosphates IA-2 and IA-2β, and islet cell antigen (ICA)-512 [17]. Type 2 diabetes, also named as adult-onset diabetes and non-insulin-dependent or insulin-resistant diabetes, produces insulin, but the body doesn’t respond to it. Ninety to ninety-five percent of diabetes patients suffer from type 2 diabetes worldwide, particularly adults above 40 years of age. The main causative factors for type 2 diabetes are inflammation, dysregulation of amylin, obesity, hypertension, increased risk of hypoglycemia, and dyslipidemia [18, 19]. Low-grade systemic inflammation and elevated levels of pro-inflammatory cytokines, i.e., IL-6 and tumor necrosis factor-α (TNF-α), which are derived from adiponectin-derived factor, are classical features of type 2 diabetes and insulin

10

G. Srivani et al.

resistance [20]. However, stimulation of the two signaling pathways, i.e., c-Jun amino N-terminal kinase (JNK) and nuclear factor kappa B (NF-κB) pathways, contributes to the development of insulin resistance and type 2 diabetes [21].

1.6

Gestational Diabetes

In pregnancy, various physiological changes occur, such as abnormal glucose levels which lead to imbalances in glucose homeostasis in the form of diabetes. About 10% of all pregnant women develop this condition, known as gestational diabetes, during the second trimester of the pregnancy [22]. Hyperinsulinemia and glucose intolerance are the hallmarks of gestational diabetes. However, enhanced insulin resistance is normally detected during pregnancy to ensure the availability of glucose for the fetus. This is mainly caused by the effects of hormones such as placental prolactin, cortisol, progesterone, placental lactogen, adiponectin, and leptin. These physiological alterations that occur during pregnancy lead to disrupted glucose tolerance, which can in turn result in gestational diabetes. Insulin contributes to various metabolic functions including increased cellular uptake of glucose, amino acids, fatty acids, and potassium ions. It is also involved in anabolic actions and enhances the cellular formation of lipids, glycogen, and proteins. Hence, these key physiological functions can be impaired if insulin action is reduced, [23], especially in the third trimester of pregnancy due to increased insulin resistance leading to decreased carbohydrate tolerance [24]. Additionally, evidence suggests that there might be an interaction between obesity, the activation of NF-κB, TNF-α, and the maternal glucose levels in gestational diabetes [25].

1.7

Pancreatic Cancer

Pancreatic cancer (PC) is the fourth most devastating cancer in the world with less than 5 years survival rate [26]. Although surgery remains the most frequent strategy for treating PC, most of the patients are diagnosed with advanced-stage disease due to the lack of proper diagnostic methods and are thus unfit for surgical treatment. Additionally, pancreatic tumors are very aggressive and develop resistance to targeted chemotherapeutic drugs [27]. Chronic pancreatitis is also one of the risk factors for pancreatic cancer, and almost 5% of chronic pancreatitis cases can develop into pancreatic cancer [28]. Both the exocrine and endocrine pancreas can develop cancer; however, most of the tumors formed by the dysregulation of acinar cells are exocrine in nature. Furthermore, 95% of pancreatic adenocarcinomas originate in epithelial cells [29, 30]. Other pancreatic malignancies include acinar cell carcinoma and mucinous cystadenocarcinoma [31]. Acinar cells that secrete digestive enzymes can differentiate into ductal cells during chronic pancreatitis and are precursors of pancreatic cancer [32, 33] (Figure 1.3). Although pancreatic cancer can result from genetic mutations in the ductal epithelial cells, these mutations, which include KRAS,

1

Biology of Pancreas and Possible Diseases

11

Pancreas

Epithelial cells (Ducts )

Activates Mutations KRAS ,BRCA2,CDKN2

Inactivates Tumor suppressor genesTP53 Abnormal cell growth Proliferation Improper Immune responses Metastasis

Fig. 1.3 Alterations in ductal epithelial cells can alter the gene regulation resulting in uncontrolled cell growth and proliferation. In pancreas, alterations in epithelial cells activate mutations (KRAS, BRCA2, CDKN2) and inactivate the tumor suppressor genes Tp53. Thus, it causes the abnormal cell growth and proliferation followed by improper immune responses associated with the occurrence and development of metastasis of pancreatic cancer

CDKN2A, BRCA2, MLH1, SMAD4, and LKB1 mutations, can activate oncogenes and inactivate tumor suppressor genes such as INK4A, p53, and ARRF. These alterations lead to uncontrolled cell growth caused by the disruption of signaling cascades, such as NF-κB, mitogen-activated protein kinase (MAPK), and PI3K– AKT pathways, that regulate cell survival and growth [31].

1.8

KRAS Mutations

PC is correlated with mutations in key oncogenes and/or tumor suppressor genes. Of these, enhanced KRAS mutations are regularly linked with tumor progression. In fact, 90% of pancreatic patients have a mutation in the exon 12th codon of the KRAS protein [34]. KRAS is activated when it binds with GTP, initiating RAF family kinase activation including RAF1, ARAF, and BRAF. Phosphorylated RAFs trigger the MEK pathway, which regulates cell growth and inhibits apoptosis, leading to pancreatic cancer progression. Furthermore, altered levels of the upstream tyrosine kinase receptors EGFR and VEGFR activate RAS [35], thus stimulating multiple signaling pathways that actively participate in tumor development, including the NF-κB, MAPK, RAL, PLC–PKC, and PI3K–AKT pathways [36, 37]. Clinically, KRAS mutation assays are used as diagnostic tools to guide treatment options for pancreatic cancer, signifying that KRAS mutations could represent prognostic biomarkers for the disease [38]. Oncogenic KRAS is inhibited by 1)

12

G. Srivani et al.

farnesyltransferase inhibitors (FTIs), which recognize the CAAX motif responsible for the transformational function of RAS downstream, the other signaling pathways [39], or 2) anti-RAS peptide vaccines at the RNA interference level in the posttranscriptional stage [38]. Recent studies have shown that TLN-4601 and salirasib, which constitute farnesyl isoprenoid small molecules, can act as PC antagonists by decreasing RAS–GTP levels and promoting cell death [40, 41].

1.9

P16/INK4/CDKN2A Mutations

Cyclin-dependent kinase (CDKN2A) (also well-known as MTS-1, P16/INK4, and CDK4I) gene is located on chromosome 9p21. Through the G1–S checkpoint, it suppresses the cell cycle development which is also necessary to regulate uncontrolled cell differentiation [42]. The CDKN2A gene is inactivated in pancreatic cancer by numerous factors such as loss of CDKN2A function, due to promoter slicing hypermethylation or homozygous deletion and loss of heterozygosity [43]. Inherited alterations in CDKN2A cause familial atypical different melanomas and greater risk of pancreatic cancer. It is one of the most commonly changed genes in cancer and the most prevalent of mutations in sporadic pancreatic cancer, with inactivation occurring in 98% of cases [43]. Some clinical studies have shown that the loss of INK4 provokes the activation of KRAS in pancreatic tumor development and promotes chemoresistance to pancreatic cancer. Hence, these mutations can act as predictive biomarkers for diagnosing the disease [38].

1.10

BRCA2 Mutations

BRCA2 is also known as FANCD1 [36]. The BRCA2 gene is positioned on chromosome no. 13Q12–13Q13 [44]. It is a multifaceted tumor suppressor and has a vital role in genetic integrity. BRCA2 exerts various functions in cell cyclespecific methods, including G2–M cell cycle checkpoint, chromatin remodeling, ubiquitylation, DNA repair, and transcriptional regulation [45]. Mitotic kinases control various functions of BRCA2 in cytokinesis, mitosis, and cell death. BRCA2 is involved in double-strand repair in the S-phase of the cell cycle, through the regulation of RAD51 recombinase [36, 46]. Inactivated and germline-mutated BRCA2 increases the risk for breast, stomach, ovarian, and pancreatic cancer [47, 48]. Furthermore, among the germline gene mutations that have been connected with greater risk of pancreatic cancer, BRCA2 mutations are the most common and identified in approximately 17% of familial pancreatic cancer (FPC) [49]. However, in the case of sporadic mutations, only a small number of cases habor somatic BRCA2 mutations [36]. Various numerical clinical studies were executed to evaluate the significance of BRCA2 as a prognostic biomarker for early diagnosis of pancreatic cancer. Huang, L. et al., 2013, carried out a study on two-stage association of pancreatic cancer, examining 981 cases and 1991 controls in the first stage followed by 2603 cases and 2877 controls. These studies revealed that the

1

Biology of Pancreas and Possible Diseases

13

c.532A > G mutagent situated in the 30 UTR is significantly correlated with sporadic pancreatic cancer (odd ratio, 1.30; CI, 95%; P < 0.0001) [50]. For instance, a pancreatic cancer patient carrying a BRCA2 617 del T mutation expresses the prolonged survival after treatment with a combination of docetaxel, gemcitabine, and capecitabine followed by single-agent irinotecan [51]. In another case, a patient with metastatic pancreatic adenocarcinoma responded to mitomycin C combination capecitabine [52], and this increased sensitivity to intercalating agents might be due to BRCA2 mutations [52]. These various cases suggest that BRCA2 mutations could be used as potential prognostic biomarkers of increased treatment sensitivity in pancreatic tumors.

1.11

SMAD4/DPC4 Mutations

The tumor suppressor gene SMAD4/DPC4 is located on chromosome no. 18q21. Fifty-five percent of pancreatic cancers harbor inactivating SMAD4/DPC4 gene mutations, which contribute to tumorigenesis by inhibiting growth suppression mechanisms and enhancing tumor progression through downregulation of the TGF-β signaling pathway [34, 53].

1.12

p53

p53 is a tumor suppressor gene, also termed as antigen NY-CO-13 [36]. It transcriptionally activates its target genes, which leads to cellular stress including DNA damage or oxidative stress and nutrient deficiency through restraining cellular functions increased by apoptosis or cell cycle arrest [54, 55]. Approximately 75% of human pancreatic cancers harbor mutations in the p53 tumor suppressor gene, thus underlining the significant role of p53 mutations or inactivation in pancreatic cancer development [56]. p53 plays key roles in stimulating cellular senescence in response to oxidative stress, hypoxia, ribonucleotide depletion, nutrient starvation, and DNA damage. Furthermore, p53 regulates several cellular functions such as metabolism, differentiation, stem cell function, motility, cell-to-cell communication, invasion, motility, and metastasis in the tumor microenvironment. These functions might be suppressed during tumor progression [57]. Recent studies show that p53 can stimulate classical DNA damage in response to cell senescence. Its target genes, associated with this response, namely, PmaiP1 (Noxa), P21 (CDKN1A), and Bbc3 (Puma), are not involved in the tumor suppressor functions [58]. Evidence from clinical studies suggests that p53 could serve as a potential prognostic biomarker for pancreatic cancer diagnosis and therapy prediction. A study performed in 57 patients with pancreatic cancer found an association between PC and p53 mutations as well as mRNA expression. Interestingly, the study results revealed that the patients with lowest p53 mRNA expression levels were associated with very deprived prognosis (p ¼ 0.032). In contrast, patients with high p53 mRNA expression were associated with significant progression (p ¼ 0.021) [59].

14

1.13

G. Srivani et al.

BRAF Mutations

The Catalogue of Somatic Mutations in Cancer (COSMIC) lists more than 30 different BRAF mutations involved in the development of various tumors including pancreatic tumors [60]. Most of these mutations are point mutations located on exon 15 in (V600E/Val 600/c.1799 T > A) the BRAF gene. The BRAF protooncogene, one of the RAF family proteins, has three isoforms, ARAF, BRAF, and RAF1 (CRAF), which are direct mediators of MAP kinase. Hence, using allosteric mechanisms, BRAF can directly stimulate MEK and extracellular signal-regulated kinase (ERK), which regulate cell growth and proliferation and control apoptosis [61, 62]. Vemurafenib (reversible ATP-competitive inhibitor) is an active inhibitor of BRAF in BRAF-mutated cancers. It is used to treat patients with metastatic melanoma and aggressive growth of a KRAS-mutant pancreatic carcinoma [63]. Therefore, the BRAF V600/E/val assay is used as a prognostic biomarker to detect various cancers, including pancreatic cancer. Ninety percent of pancreatic tumors carrying mutations in KRAS stimulate several downstream signaling pathways, such as RAS–RAF–MEK–ERK (MAPK) and PI3K–AKT–mTOR pathways.

1.14

RAS–RAF–MEK–ERK (MAPK) Signaling Pathway

The RAS–MEK signaling pathways play a major role in cellular activities such as cell division, growth, differentiation, proliferation, progression, survival, and metabolism in response to upstream signals. The dysfunction of intermediates of the MAPK pathway results in diverse effects such as tumorigenesis and the ability to function without mitogenic signals, abnormal cell proliferation, and the inhibition the apoptosis. Hence, alterations and mutations in MAPK signaling promote tumor progression [64]. BRAF is the initiator of the MAPK pathway. The extracellular signal-regulated kinase (ERK)–mitogen-activated protein kinase (MAPK) cascade is characterized by sequential molecule activation, initiated by the small GTPase and different protein kinases that phosphorylate MAPKKK then MAPKK, which functions as a gate keeper and phosphorylates/stimulates the effector kinase–mitogen-activated protein kinase (MAPK). To date, the MAPK cascade has been divided in to six groups in mammals, namely, ERK1/2, ERK 3/4, ERK 5, ERK 7/8, JNK 1/2/3 (c-Jun amino N-terminal kinase), and p38 isoforms [65, 66]. The MAPK signaling pathway has various effects which can control cell cycle progression, differentiation, and cell senescence [67]. A review of literature documents the complexity of this pathway, revealing the various kinases, apoptotic regulators, caspase executioner families, and transcription factors, which can be either activated or inactivated through phosphorylation of proteins. Moreover, the MAPK pathway stimulates the transcription of multiple genes. For instance, RAF can stimulate the phosphorylation of protein involved in cell death either by downregulating MEK and ERK or self-regulating MEK and ERK.

1

Biology of Pancreas and Possible Diseases

15

The irregular activation of MAPK signaling can lead to the development of cancers via various mechanisms, especially genetic mutations. These mutations can occur due to the upregulation of membrane receptors and BRAF and RAS genes as well as other genes involved in various pathways, viz., PTEN, PI3K, and Akt, which are essential for regulating RAF activity [68, 69]. RAS has been recognized as an oncogene that is mutationally activated in 25% of all the cancers. It is associated with either by downregulation of MEK and ERK or self-regulation of MEK and ERK and eminent incidence in colon (50%), lung (30%), pancreas (90%), melanoma (25%), and thyroid (50%) cancers [67, 70]. In view of these key functions, the RAS/RAF/MEK/ERK pathway is an essential pathway for remedial interventions. Accordingly, various inhibitors are currently being developed to target the MEK cascade and exert effective antiproliferative activity. For instance, CI-1040 represents the first MEK inhibitor and succeed in in vivo and in vitro clinical phases, although it did not demonstrate potent antitumor activity in the phase II clinical studies [71], indicating that further pharmacokinetic enhancements are required to improve the efficacy of these inhibitors.

1.15

PI3K–AKT–mTOR Signaling Pathway

1.15.1 PI3K Signaling Since PI3K innovation in the 1980s, the family of lipid kinases named phosphatidylinositol 3-kinases (PI3Ks) has gathered wide interest. It has been found to contribute to various cellular processes (cell growth, proliferation, migration, and survival) [72]. PI3K signaling is one of the most commonly dysregulated pathways in cancers [73]. PI3K is activated by G-protein-coupled receptors (GPCRs) and receptor tyrosine kinases (RTKs). This leads to the PI3K autophosphorylation, which results in the activation of the serine/threonine kinase Akt pathway and other downstream effectors [72, 74]. Abnormal activation of this pathway affects various cellular processes, such as metabolism, cell cycle progression, survival, regulation of cell death, genomic instability, and protein synthesis [75]. Various agents have been developed to target the PI3K–Akt pathway, among which LY294002 and wortmannin act as potential inhibitors of the PI3K pathway [75]. The PI3K pathway inhibits cell death and regulates cell differentiation during tumor development. PI3K activates the phosphatidylinositol 3,4,5-triphosphate through the phosphorylation of phosphatidylinositol 4,5-bisphosphate leading to the stimulation of the AKT pathway, which targets various downstream effectors [76]. mTOR is a major target of this pathways. It is crucial for cellular growth and metabolism and is involved in the activation of the initiation factor 4E-binding protein (4EBP1) in eukaryotic translation [77].

16

G. Srivani et al.

AKT signaling IGFIR PDGFR EGFR VEGFR

Growth Factors

PI3K

PIP2 PIP3 AKT

mTOR

FoXo

4EBP1

Cdk

eIF4E HIF1α

GERS/GAPS

Rac/Rho family Inhibits apoptosis Cell cycle arrest

Survival , Invasion, Proliferation , Metabolism

Cell cycle progression Cyto skeleton motility NADPH oxidation

Fig. 1.4 AKT signaling regulates several downstream substrates that are participated in the various cellular events. Growth factors (IGFIR, PDGFRS, EGFR, and VEGFR) activates the PI3K that phosphorylates and activates the AKT. Thus, activated AKT downregulates the multiple targets (FOXO, IKK, NF-κB) in order to inhibit the apoptosis and increase the cell proliferation and survival. Directly or indirectly, AKT activates the mTOR that promotes the inactivation of HIF-1α which results to cell proliferation and invasion. AKT activates the Rac/Rho family and promotes the cytoskeleton motility and NADPH oxidation

1.15.2 AKT Signaling AKT (protein kinase B) is an essential regulator of human physiology. It plays a significant role in promoting several cellular functions required for cell growth, proliferation, survival, and metabolism [78]. The AKT signaling cascade (Fig. 1.4) leads to the activation of various transcription factors including HIF-1α, NF-κB, and FOXO which play a vital role in cell cycle regulation and the suppression of cell death [79].AKT signaling alterations cause several genetic abnormalities in different types of tumors. These genetic abnormalities underline the key contribution of AKT in the development of pancreatic cancer.

1.15.3 mTOR Signaling Mammalian target of rapamycin (mTOR) is a serine/threonine protein kinase and one of the downstream targets of the PI3K signaling pathway (Fig. 1.5). It exits as

1

Biology of Pancreas and Possible Diseases

Energy

17 Nutrients Amino acids

Oxygen /stress

Growth factors

mTOR

Ribosome biogenesis

4EbP1

S6K1

Regulation of Transcription

Cell Growth Cell proliferation

Fig. 1.5 Functions of the mTOR signaling. Mammalian target of rapamycin (mTOR) is a serine/ threonine protein kinase. Oxygen levels, growth factors, nutrients, and intracellular energy activate the mTOR and thus regulate the anabolic and catabolic processes; downregulate the several targets which include ribosome biogenesis, 4EBP1 (eukaryotic initiation factor 4E-binding protein), and S6K1 (ribosomal S6 protein kinase 1); and result in the regulation of transcription followed by cell growth, proliferation, and cell metabolism

two functional catalytic subunits formed by mTORC1 and mTORC2. mTOR contributes to several growth-related functions including enhanced cell growth, proliferation and survival, rearrangement of actin cytoskeleton, and protein degradation [80]. mTORC1 is composed of mTOR, Gbl, and Raptor, which act as a scaffold protein interlinked to the mTOR kinase, which activates the mTORC1 signaling. Following PI3K/AKT signaling, mTORC1 stimulates anabolic processes and suppresses catabolic reactions in the cell. Several signaling molecules (growth factors, energy, nutrients, and amino acids) activate mTORC1, thus promoting diverse cellular processes, i.e., cell growth, cell differentiation, migration, and cell metabolism. On the other hand, inadequate levels of these molecules can alter mTORC1 function and promote macroautophagy. mTORC2, on the other hand, is composed of mTOR, Gbl, and rapamycin-insensitive companion of mTOR (Rictor) and participates in AKT signaling. It contributes to cell differentiation, and its main function is to regulate the actin cytoskeleton. The stimulation of the mTOR signaling cascade modulates various cellular processes including the translation of several proteins, such as HIF-1α, which activates VEGF expression in angiogenesis, and cyclin D1, which stimulates cell cycle development [81, 82]. mTORC1 is extremely susceptible to rapamycin, while mTORC2 is less sensitive to it. Driscoll et al. [82] revealed that the suppression of mTORC2/PI3K signaling is a possible therapeutic strategy for pancreatic cancer (Fig. 1.6.) The above survival signaling PI3K/AKT/mTOR cascade is fully involved in the regulation of cell proliferation and apoptosis, and more than 50% of cancers are

18

G. Srivani et al.

Growth factors

PI3K

Gbl

mTORC2 Rictor AKT

TSC Nutrients Gbl

Cell survival Metabolism Proliferaon

mTORC1 Raptor

S6K

4EBP

HIF1α

Survival , Angiogenesis

Ribosome biogenesis

Fig. 1.6 Activities of mTORC1/2: mTOR is the catalytic components of the two subunits named as mTORC1 and mTORC2. mTORC1 is composed of mTOR, Gbl, and Raptor; mTORC2 is composed of mTOR, Gbl, and Rictor. Growth factors activate mTORC2 by activation of PI3K, and activated mTORC2 stimulates the AKT, thus regulating the various cellular functions such as cell growth, proliferation, survival, and metabolism. AKT activates mTORC1 by inhibiting the tuberous sclerosis complex (TSC) which leads to the activation of downstream targets S6K, 4EBP, and HIF-1α and promotes the ribosome biogenesis, cell survival, and angiogenesis

caused by alterations in this pathway, including pancreatic cancer [83]. The overexpression of AKT and its involvement in the cellular plasticity of the pancreas demonstrate how AKT signaling could play a crucial role in the development of pancreatic tumors [84]. Therefore, the PI3K–AKT pathway is crucial to the progression of pancreatic cancer. HIF-1α is a transcription factor and a key regulator of angiogenesis, necessary for the cellular response to low oxygen. The PI3K/AKT/mTOR pathway increases HIF-1α gene expression, thus allowing tumor cells to survive and develop under hypoxic conditions. Therefore, mTOR and HIF-1α could serve as prognostic targets for drug development [85]. Furthermore, Peng T and Dou QP showed in their recent study that the mTOR inhibitor everolimus can enhance the efficacy of gemcitabine chemotherapy in resistant pancreatic tumors. This indicates that mTOR inhibitors with improved efficacies are required to improve antitumor activity and increase the survival rate in pancreatic cancer patients [86]. In a phase II clinical study of patients treated with a combination of capecitabine and everolimus, Kordes S et al. observed moderate antitumor activity and minimal

1

Biology of Pancreas and Possible Diseases

19

toxicity levels, thus signifying that combination therapies could be effective for treating pancreatic cancer [87]. One of the undergoing clinical phase I/II studies is investigating the sorafenib combination everolimus in patients with advanced pancreatic cancer (NCT00981162).

1.16

Endocrine Pancreas

Pancreatic endocrine tumors occur very rarely, with roughly 5 per in 1,000,000/year. Pancreatic endocrine tumors (PETs) are often termed as “islet cell tumors” or “pancreatic neuroendocrine tumors” (PNETs) [88]. PETs originate from the endocrine (hormonal) cells and nervous system within the pancreas [89] and are mostly associated with the lungs and gastrointestinal tract and also found in the pancreas [89, 90]. However, PETs have been shown to decelerate growing and are not as much aggressive as invasive ductal carcinomas of the pancreas. However, when they reach the metastatic stage, they become life-threatening and more difficult to cure [91]. The US Surveillance, Epidemiology, and End Results (SEER) survey from the year 1973–2014 indicates that PETs amount to 3.6% of all neuroendocrine tumors (NETs) [91]. PETs are a very rare subgroup of pancreatic tumors and represent 1–2% of all the pancreatic cancers [92]. The incidence of NETs has significantly improved, and the prevalence of PETs has also drastically increased in each primary tumor site [93]. Furthermore, about 20% of PETs are nonfunctional. Functional tumors might be benign or malignant in nature, with no environmental factors involved in tumor development [88]. Pancreatic NETs are classified as functional NETs and nonfunctional NETs. Functional NETs are associated with clinical symptoms and associated to the type of hormones released including gastrin, insulin, glucagon, somatostatin, and vasoactive intestinal peptide, whereas nonfunctional NETs are not associated with clinical symptoms either because no hormone is secreted or the other substances (subunits of human chorionic gonadotropin, ghrelin, neurotensin, neuron-specific enolase, chromogranins) that are secreted do not display clinical symptoms [91]. Common functional syndromes of this disease are gastrinomas, insulinoma, glucagonomas, VIPomas, and somatostatinoma, caused by the secretion of gastrin, insulin, vasoactive intestinal peptide, glucagon, and somatostatin [94]. PETs are non-inherited (sporadic) and approximately 10–30% of nonfunctional PETs display an endocrine tumor syndrome which includes tuberous fibromatosis (TSC), multiple endocrine neoplasia syndrome type 1(MEN 1), Von Recklinghausen’s syndrome (neurofibromatosis), and von Hippel–Lindau syndrome [95, 96]. The nonfunctional familial syndromes occur due to inherited germline loss and gain of functional mutations in tumor suppressor genes [96].

20

1.17

G. Srivani et al.

Conclusion

The pancreas plays a critical role and various physiological functions and is associated with various deadly diseases (pancreatitis, diabetes, and pancreatic cancer), caused by improper immune responses associated with genetic mutations and lifestyle modifications. Recently, several studies elucidate that alcohol, smoking, and obesity are the main causative risk factors for development of pancreatitis to pancreatic cancer. Predominately, abnormal zymogen activation and genetic mutations particularly abnormal function of trypsinogen lead to acute pancreatitis and chronic pancreatitis. Furthermore, long-standing chronic pancreatitis is the development of diabetes mellitus. Moreover, activation of oncogenes and inactivated tumor suppressor genes contributes to the development and progression of chronic pancreatitis to pancreatic cancer. Among these disorders, pancreatic cancer has a very poor prognosis and survival. The lack of early detection methods reduces the chances of identifying the tumor at an early stage. Therefore, researchers are trying to develop enhanced diagnostic tools to aid in preventing and treating the disease to increase survival outcomes. Recently, the improved understanding of the molecular mechanisms mediating PC development has resulted in the design of enhanced targeted therapies including combination therapies and photochemical therapies that target numerous signaling pathways and prevent tumor reoccurrence.

References 1. Longnecker DS (2014) Anatomy and histology of the pancreas. Pancreapedia, The Exocrine Pancreas Knowledge Base 2. Cappell MS (2008) Acute pancreatitis: etiology, clinical presentation, diagnosis, and therapy. Med Clin N Am 92(4):889–923 3. Ahmed A, Azim A, Gurjar M, Baronia AK (2016) Hypocalcemia in acute pancreatitis revisited. Indian J Crit Care Med: Peer-Rev Off Publ Indian Soc Crit Care Med 20(3):173–177 4. Whitcomb DC (1999) The spectrum of complications of hereditary pancreatitis: is this a model for future gene therapy? Gastroenterol Clin 28(3):525–541 5. Shelton CA, Whitcomb DC (2016) Hereditary pancreatitis. Pancreapedia, The Exocrine Pancreas Knowledge Base 6. Zyromski NJ, Mathur A, Pitt HA, Lu D, Gripe JT, Walker JJ, Yancey K, Wade TE, SwartzBasile DA (2008) A murine model of obesity implicates the adipokine milieu in the pathogenesis of severe acute pancreatitis. Am J Physiol Gastrointest Liver Physiol 295(3):G552–G558 7. Johnson AR, Milner JJ, Makowski L (2012) The inflammation highway: metabolism accelerates inflammatory traffic in obesity. Immunol Rev 249(1):218–238 8. Kolodecik T, Shugrue C, Ashat M, Thrower EC (2014) Risk factors for pancreatic cancer: underlying mechanisms and potential targets. Front Physiol 4:415 9. Hoque R, Mehal WZ (2015) Inflammasomes in pancreatic physiology and disease. Am J Physiol Gastrointest Liver Physiol 308(8):G643–G651 10. Habtezion A (2015) Inflammation in acute and chronic pancreatitis. Curr Opin Gastroenterol 31 (5):395 11. Ewald N, Hardt PD (2013) Diagnosis and treatment of diabetes mellitus in chronic pancreatitis. World J Gastroenterol: WJG 19(42):7276

1

Biology of Pancreas and Possible Diseases

21

12. Piero M, Nzaro G, Njagi J (2015) Diabetes mellitus? a devastating metabolic disorder. Asian J Biomed Pharm Sci 4(40) 13. Ogurtsova K, da Rocha FJ, Huang Y, Linnenkamp U, Guariguata L, Cho N, Cavan D, Shaw J, Makaroff L (2017) IDF diabetes atlas: global estimates for the prevalence of diabetes for 2015 and 2040. Diabetes Res Clin Pract 128:40–50 14. Karalliedde J, Gnudi L (2014) Diabetes mellitus, a complex and heterogeneous disease, and the role of insulin resistance as a determinant of diabetic kidney disease. Nephrol Dial Transplant 31(2):206–213 15. Sicree R, Shaw J, Zimmet P (2006) Prevalence and projections. Diabetes atlas 3:16–104 16. Coppieters KT, Dotta F, Amirian N, Campbell PD, Kay TW, Atkinson MA, Roep BO, von Herrath MG (2012) Demonstration of islet-autoreactive CD8 T cells in insulitic lesions from recent onset and long-term type 1 diabetes patients. J Exp Med 209(1):51–60. jem. 20111187 17. Hassan GA, Sliem HA, Ellethy AT, Salama ME-S (2012) Role of immune system modulation in prevention of type 1 diabetes mellitus. Indian journal of endocrinology and metabolism 16 (6):904 18. Boden G, Laakso M (2004) Lipids and glucose in type 2 diabetes: what is the cause and effect? Diabetes Care 27(9):2253–2259 19. Hieronymus L, Griffin S (2015) Role of amylin in type 1 and type 2 diabetes. Diabetes Educ 41 (1_suppl):47S–56S 20. Andreasen AS, Kelly M, Berg RMG, Møller K, Pedersen BK (2011) Type 2 diabetes is associated with altered NF-κB DNA binding activity, JNK phosphorylation, and AMPK phosphorylation in skeletal muscle after LPS. PLoS One 6(9):e23999–e23999 21. Kahn SE, Hull RL, Utzschneider KM (2006) Mechanisms linking obesity to insulin resistance and type 2 diabetes. Nature 444(7121):840 22. Öztekin Ö (2007) New insights into the pathophysiology of gestational diabetes mellitus: possible role of human leukocyte antigen-G. Med Hypotheses 69(3):526–530 23. Al-Noaemi MC, Shalayel MHF (2011) Pathophysiology of gestational diabetes mellitus: the past, the present and the future. In: Gestational Diabetes. InTech 24. Shalayel MHF, Al-Noaemi MC, Ahmed SA (2011) Insulin resistance in the third trimester of pregnancy suffering from gestational diabetes mellitus or impaired glucose tolerance. In: Gestational Diabetes. InTech 25. Lappas M, Hiden U, Desoye G, Froehlich J, Mouzon SHD, Jawerbaum A (2011) The role of oxidative stress in the pathophysiology of gestational diabetes mellitus. Antioxid Redox Signal 15(12):3061–3100 26. Miller KD, Siegel RL, Lin CC, Mariotto AB, Kramer JL, Rowland JH, Stein KD, Alteri R, Jemal A (2016) Cancer treatment and survivorship statistics, 2016. CA Cancer J Clin 66 (4):271–289 27. McCarroll JA, Naim S, Sharbeen G, Lee J, Kavallaris M, Goldstein D, Phillips PA (2014) Role of pancreatic stellate cells in chemoresistance in pancreatic cancer. Front Physiol 5:141 28. Pinho AV, Chantrill L, Rooman I (2014) Chronic pancreatitis: a path to pancreatic cancer. Cancer Lett 345(2):203–209 29. Hezel AF, Kimmelman AC, Stanger BZ, Bardeesy N, DePinho RA (2006) Genetics and biology of pancreatic ductal adenocarcinoma. Genes Dev 20(10):1218–1249 30. Hidalgo M, Cascinu S, Kleeff J, Labianca R, Löhr J-M, Neoptolemos J, Real FX, Van Laethem J-L, Heinemann V (2015) Addressing the challenges of pancreatic cancer: future directions for improving outcomes. Pancreatology 15(1):8–18 31. Bardeesy N, DePinho RA (2002) Pancreatic cancer biology and genetics. Nat Rev Cancer 2 (12):897 32. Becker AE, Hernandez YG, Frucht H, Lucas AL (2014) Pancreatic ductal adenocarcinoma: risk factors, screening, and early detection. World J Gastroenterol: WJG 20(32):11182 33. Pelosi E, Castelli G, Testa U (2017) Pancreatic Cancer: molecular characterization, clonal evolution and Cancer stem cells. Biomedicine 5(4):65

22

G. Srivani et al.

34. Feldmann G, Maitra A (2008) Molecular genetics of pancreatic ductal adenocarcinomas and recent implications for translational efforts. J Mol Diagn 10(2):111–122 35. Shawver LK, Slamon D, Ullrich A (2002) Smart drugs: tyrosine kinase inhibitors in cancer therapy. Cancer Cell 1(2):117–123 36. Cicenas J, Kvederaviciute K, Meskinyte I, Meskinyte-Kausiliene E, Skeberdyte A (2017) KRAS, TP53, CDKN2A, SMAD4, BRCA1, and BRCA2 mutations in pancreatic cancer. Cancers 9(5):42 37. McCormick F (2015) KRAS as a therapeutic target. Clin Cancer Res 21(8):1797–1801 38. Bournet B, Buscail C, Muscari F, Cordelier P, Buscail L (2016) Targeting KRAS for diagnosis, prognosis, and treatment of pancreatic cancer: hopes and realities. Eur J Cancer 54:75–83 39. Baines AT, Xu D, Der CJ (2011) Inhibition of Ras for cancer treatment: the search continues. Future Med Chem 3(14):1787–1808 40. Campbell PM, Boufaied N, Fiordalisi JJ, Cox AD, Falardeau P, Der CJ, Gourdeau H (2010) TLN-4601 suppresses growth and induces apoptosis of pancreatic carcinoma cells through inhibition of Ras-ERK MAPK signaling. J Mol Signal 5(1):18 41. Kloog Y, Blum R, Cox AD (2008) Inhibitors of chronically active Ras: potential for treatment of human malignancies. Recent Pat Anticancer Drug Discov 3(1):31–47 42. Agarwal P, Kabir FML, De Innocentes P, Bird RC (2012) Tumor suppressor gene p16/INK4A/ CDKN2A and its role in cell cycle exit, differentiation, and determination of cell fate. In: Tumor suppressor genes. InTech 43. Schutte M, Hruban RH, Geradts J, Maynard R, Hilgers W, Rabindran SK, Moskaluk CA, Hahn SA, Schwarte-Waldhoff I, Schmiegel W (1997) Abrogation of the Rb/p16 tumor-suppressive pathway in virtually all pancreatic carcinomas. Cancer Res 57(15):3126–3130 44. Hamann U, Liu X, Lange S, Ulmer H, Benner A, Scott R (2002) Contribution of BRCA2 germline mutations to hereditary breast/ovarian cancer in Germany. J Med Genet 39(3):e12–e12 45. Rudkin TM, Foulkes WD (2005) BRCA2: breaks, mistakes and failed separations. Trends Mol Med 11(4):145–148 46. Lee H (2014) Cycling with BRCA2 from DNA repair to mitosis. Exp Cell Res 329(1):78–84 47. Welcsh PL, King M-C (2001) BRCA1 and BRCA2 and the genetics of breast and ovarian cancer. Hum Mol Genet 10(7):705–713 48. Friedenson B (2005) BRCA1 and BRCA2 pathways and the risk of cancers other than breast or ovarian. Medscape Gen Med 7(2):60 49. Carnevale J, Ashworth A (2015) Assessing the significance of BRCA1 and BRCA2 mutations in pancreatic cancer. Am Soc Clin Oncol 33:3080 50. Huang L, Wu C, Yu D, Wang C, Che X, Miao X, Zhai K, Chang J, Jiang G, Yang X (2013) Identification of common variants in BRCA2 and MAP 2K4 for susceptibility to sporadic pancreatic cancer. Carcinogenesis 34(5):1001–1005 51. James E, Waldron-Lynch MG, Saif MW (2009) Prolonged survival in a patient with BRCA2 associated metastatic pancreatic cancer after exposure to camptothecin: a case report and review of literature. Anti-Cancer Drugs 20(7):634–638 52. Chalasani P, Kurtin S, Dragovich T (2008) Response to a third-line mitomycin C (MMC)-based chemotherapy in a patient with metastatic pancreatic adenocarcinoma carrying germline BRCA2 mutation. JOP 9(3):305–308 53. Cao D, Ashfaq R, Goggins MG, Hruban RH, Kern SE, Iacobuzio-Donahue CA (2008) Differential expression of multiple genes in association with MADH4/DPC4/SMAD4 inactivation in pancreatic cancer. Int J Clin Exp Pathol 1(6):510 54. Brady CA, Jiang D, Mello SS, Johnson TM, Jarvis LA, Kozak MM, Broz DK, Basak S, Park EJ, McLaughlin ME (2011) Distinct p53 transcriptional programs dictate acute DNA-damage responses and tumor suppression. Cell 145(4):571–583 55. Levy N, Yonish-Rouach E, Oren M, Kimchi A (1993) Complementation by wild-type p53 of interleukin-6 effects on M1 cells: induction of cell cycle exit and cooperativity with c-myc suppression. Mol Cell Biol 13(12):7942–7952

1

Biology of Pancreas and Possible Diseases

23

56. Mello SS, Valente LJ, Raj N, Seoane JA, Flowers BM, McClendon J, Bieging-Rolett KT, Lee J, Ivanochko D, Kozak MM (2017) A p53 super-tumor suppressor reveals a tumor suppressive p53-Ptpn14-Yap axis in pancreatic cancer. Cancer Cell 32(4):460–473. e466 57. Bieging KT, Mello SS, Attardi LD (2014) Unravelling mechanisms of p53-mediated tumour suppression. Nat Rev Cancer 14(5):359–370 58. Valente Liz J, Gray Daniel HD, Michalak Ewa M, Pinon-Hofbauer J, Egle A, Scott Clare L, Janic A, Strasser A (2013) p53 efficiently suppresses tumor development in the complete absence of its cell-cycle inhibitory and Proapoptotic effectors p21, Puma, and Noxa. Cell Rep 3(5):1339–1345 59. Grochola LF, Taubert H, Greither T, Bhanot U, Udelnow A, Würl P (2011) Elevated transcript levels from the MDM2 P1 promoter and low p53 transcript levels are associated with poor prognosis in human pancreatic ductal adenocarcinoma. Pancreas 40(2):265–270 60. Bamford S, Dawson E, Forbes S, Clements J, Pettett R, Dogan A, Flanagan A, Teague J, Futreal PA, Stratton MR (2004) The COSMIC (Catalogue of Somatic Mutations in Cancer) database and website. Br J Cancer 91(2):355 61. Davies H, Bignell GR, Cox C, Stephens P, Edkins S, Clegg S, Teague J, Woffendin H, Garnett MJ, Bottomley W (2002) Mutations of the BRAF gene in human cancer. Nature 417(6892):949 62. Wan PT, Garnett MJ, Roe SM, Lee S, Niculescu-Duvaz D, Good VM, Project CG, Jones CM, Marshall CJ, Springer CJ (2004) Mechanism of activation of the RAF-ERK signaling pathway by oncogenic mutations of B-RAF. Cell 116(6):855–867 63. Grey A, Cooper A, McNeil C, O’toole S, Thompson J, Grimison P (2014) Progression of KRAS mutant pancreatic adenocarcinoma during vemurafenib treatment in a patient with metastatic melanoma. Intern Med J 44(6):597–600 64. Cargnello M, Roux PP (2011) Activation and function of the MAPKs and their substrates, the MAPK-activated protein kinases. Microbiol Mol Biol Rev 75(1):50–83 65. Krens SG, Spaink HP, Snaar-Jagalska BE (2006) Functions of the MAPK family in vertebratedevelopment. FEBS Lett 580(21):4984–4990 66. Kyriakis JM, Avruch J (2001) Mammalian mitogen-activated protein kinase signal transduction pathways activated by stress and inflammation. Physiol Rev 81(2):807–869 67. Liu F, Yang X, Geng M, Huang M (2018) Targeting ERK, an Achilles’ Heel of the MAPK pathway, in cancer therapy. Acta Pharm Sin B 8(4):552–562 68. McCubrey JA, Steelman LS, Chappell WH, Abrams SL, Wong EWT, Chang F, Lehmann B, Terrian DM, Milella M, Tafuri A et al (2007) Roles of the Raf/MEK/ERK pathway in cell growth, malignant transformation and drug resistance. Biochim Biophys Acta 1773 (8):1263–1284 69. Garnett MJ, Marais R (2004) Guilty as charged: B-RAF is a human oncogene. Cancer Cell 6 (4):313–319 70. Hobbs GA, Der CJ, Rossman KL (2016) RAS isoforms and mutations in cancer at a glance. J Cell Sci 129(7):1287–1292 71. Allen LF, Sebolt-Leopold J, Meyer MB (2003) CI-1040 (PD184352), a targeted signal transduction inhibitor of MEK (MAPKK). In: Seminars in oncology: 2003. Elsevier, pp 105–116 72. Liu P, Cheng H, Roberts TM, Zhao JJ (2009) Targeting the phosphoinositide 3-kinase pathway in cancer. Nat Rev Drug Discov 8(8):627–644 73. Baer R, Cintas C, Therville N, Guillermet-Guibert J (2015) Implication of PI3K/Akt pathway in pancreatic cancer: when PI3K isoforms matter? Adv Biol Reg 59:19–35 74. Murthy D, Attri KS, Singh PK (2018) Phosphoinositide 3-kinase signaling pathway in pancreatic ductal adenocarcinoma progression, pathogenesis, and therapeutics. Front Physiol 9:335–335 75. Ebrahimi S, Hosseini M, Shahidsales S, Maftouh M, Ferns GA, Ghayour-Mobarhan M, Mahdi Hassanian S, Avan A (2017) Targeting the Akt/PI3K signaling pathway as a potential therapeutic strategy for the treatment of pancreatic cancer. Curr Med Chem 24(13):1321–1331

24

G. Srivani et al.

76. Cantley LC, Neel BG (1999) New insights into tumor suppression: PTEN suppresses tumor formation by restraining the phosphoinositide 3-kinase/AKT pathway. Proc Natl Acad Sci 96 (8):4240–4245 77. Payne S, Maher M, Tran N, Van De Hey D, Foley T, Yueh A, Leystra A, Pasch C, Jeffrey J, Clipson L (2015) PIK3CA mutations can initiate pancreatic tumorigenesis and are targetable with PI3K inhibitors. Oncogene 4(10):e169 78. Gonzalez E, McGraw TE (2009) The Akt kinases: isoform specificity in metabolism and cancer. Cell Cycle 8(16):2502–2508 79. Zhang Y, Ng PK-S, Kucherlapati M, Chen F, Liu Y, Tsang YH, de Velasco G, Jeong KJ, Akbani R, Hadjipanayis A (2017) A pan-cancer proteogenomic atlas of PI3K/AKT/mTOR pathway alterations. Cancer Cell 31(6):820–832. e823 80. Foster KG, Fingar DC (2010) Mammalian target of rapamycin (mTOR): conducting the cellular signaling symphony. J Biol Chem 285(19):14071–14077 81. Hay N, Sonenberg N (2004) Upstream and downstream of mTOR. Genes Dev 18 (16):1926–1945 82. Driscoll DR, Karim SA, Sano M, Gay DM, Jacob W, Yu J, Mizukami Y, Gopinathan A, Jodrell DI, Evans TJ (2016) mTORC2 signaling drives the development and progression of pancreatic cancer. Cancer Res 76:6911 83. Falasca M, Selvaggi F, Buus R, Sulpizio S, Edling EC (2011) Targeting phosphoinositide 3-kinase pathways in pancreatic cancer-from molecular signalling to clinical trials. Anti-Cancer Agents Med Chem (Formerly Current Medicinal Chemistry-Anti-Cancer Agents) 11 (5):455–463 84. Elghazi L, Weiss AJ, Barker DJ, Callaghan J, Staloch L, Sandgren EP, Gannon M, Adsay VN, Bernal-Mizrachi E (2009) Regulation of pancreas plasticity and malignant transformation by Akt signaling. Gastroenterology 136(3):1091–1103. e1098 85. Harris AL (2002) Hypoxia—a key regulatory factor in tumour growth. Nat Rev Cancer 2(1):38 86. Peng T, Dou QP (2017) Everolimus inhibits growth of gemcitabine-resistant pancreatic Cancer cells via induction of caspase-dependent apoptosis and G2/M arrest. J Cell Biochem 118 (9):2722–2730 87. Kordes S, Klümpen H-J, Weterman MJ, Schellens JH, Richel DJ, Wilmink JW (2015) Phase II study of capecitabine and the oral mTOR inhibitor everolimus in patients with advanced pancreatic cancer. Cancer Chemother Pharmacol 75(6):1135–1141 88. Brunicardi FC, Anderson D, Billiar TR, Dunn DL, Hunter JG, Pollock RE (2006) Schwartz’Principles of surgery: self-assessment and board review. McGraw Hill Professional, New York 89. Burns WR, Edil BH (2012) Neuroendocrine pancreatic tumors: guidelines for management and update. Curr Treat Options in Oncol 13(1):24–34 90. Yao JC, Hassan M, Phan A, Dagohoy C, Leary C, Mares JE, Abdalla EK, Fleming JB, Vauthey J-N, Rashid A (2008) One hundred years after “carcinoid”: epidemiology of and prognostic factors for neuroendocrine tumors in 35,825 cases in the United States. J Clin Oncol 26 (18):3063–3072 91. Kondo NI, Ikeda Y (2014) Practical management and treatment of pancreatic neuroendocrine tumors. Gland Surg 3(4):276–283 92. Halfdanarson TR, Rabe KG, Rubin J, Petersen GM (2008) Pancreatic neuroendocrine tumors (PNETs): incidence, prognosis and recent trend toward improved survival. Ann Oncol Off J Eur Soc Med Oncol 19(10):1727–1733 93. Modlin IM, Oberg K, Chung DC, Jensen RT, de Herder WW, Thakker RV, Caplin M, Delle Fave G, Kaltsas GA, Krenning EP (2008) Gastroenteropancreatic neuroendocrine tumours. Lancet Oncol 9(1):61–72 94. Metz DC, Jensen RT (2008) Gastrointestinal neuroendocrine tumors: pancreatic endocrine tumors. Gastroenterology 135(5):1469–1492

1

Biology of Pancreas and Possible Diseases

25

95. Bartolini I, Bencini L, Risaliti M, Ringressi MN, Moraldi L, Taddei A (2018) Current Management of Pancreatic Neuroendocrine Tumors: from Demolitive surgery to observation. Gastroenterol Res Pract 2018:1 96. Ro C, Chai W, Yu VE, Yu R (2013) Pancreatic neuroendocrine tumors: biology, diagnosis, and treatment. Chin J Cancer 32(6):312–324

2

Pancreatitis: Clinical Aspects of Inflammatory Phenotypes Nyshadham S. N. Chaitanya and Aramati BM Reddy

Abstract

Pancreatitis is a condition in which the pancreas becomes inflamed and gets damaged when the digestive enzymes are activated before they are released into the small intestine and attacks the pancreas. When these activated digestive enzymes are activated still in the pancreas, irritating the cells of our pancreas causes inflammation. These cause complications such as the formation of the pseudocyst, breathing problem, failure in the kidney, diabetes and pancreatic cancer. Two forms of pancreatitis include acute and chronic. Acute pancreatitis is a condition in which inflammation lasts for short time, while chronic pancreatitis is a condition in which inflammation lasts for a longer time. Gallstones and the gallbladder in the pancreas can be removed by surgical treatment. The probability of developing pancreatitis can be reduced by alcohol cessation and preventing gallstone complication. Risk factors include hereditary, hyperlipidaemia, smok ing, cystic fibrosis and usage of certain medicines such as oestrogen and tetracy cline. The tests to detect pancreatitis include enzymes of the pancreas, liver and kidney. Signs of infection include fever/fatigue, anaemia and decreased electrolyte and calcium level. In acute pancreatitis, the patient’s diet consists of bowel rest, and for chronic pancreatitis, it includes low-fat diet and high carbohydrates. Keywords

Acinar cell injury · Acute pancreatitis · Chronic pancreatitis · Secretin-enhanced MR cholangiography · Fluid therapy

N. S. N. Chaitanya · A. BM Reddy (*) Laboratory of Cell Signalling, Department of Animal Biology, School of Life Sciences, University of Hyderabad, Hyderabad, Telangana, India e-mail: [email protected] # Springer Nature Singapore Pte Ltd. 2019 G. P. Nagaraju, A. BM Reddy (eds.), Exploring Pancreatic Metabolism and Malignancy, https://doi.org/10.1007/978-981-32-9393-9_2

27

28

N. S. N. Chaitanya and A. BM Reddy

Abbreviations AP CP CRP ERCP FAIMS RDW VOC

2.1

Acute pancreatitis Chronic pancreatitis C-reactive protein Endoscopic retrograde cholangiopancreatography Field asymmetric waveform ion mobility spectrometry Red cell distribution width Volatile organic compounds

Introduction

Acute pancreatitis accounts for approximately 56 cases per one-tenth million persons per year in the United Kingdom, while in the United States over 220 thousand are hospitalised per year. Formation of gallstones is the common cause for acute pancreatitis, about 25 per hundred of acute pancreatitis cases were due to alcohol. Gallstones of patients smaller than 5 mm in size are thought to have higher risk of gallstone pancreatitis. Obstruction at sphincter of Oddi leads to injured pancreatic ducts. Acute pancreatitis includes improper release and activation of pancreatic enzymes, and trypsin plays a role in activating these proenzymes. Its inappropriate activation leads to pancreatic inflammation and triggers the release of IL-1, IL-6 and IL-8; TNF-α; and platelet-activating factor [24]. Advancement in magnetic resonance imaging had allowed the identification and characterisation of pancreatic disorders [25]. Chronic pancreatitis leads to pancreatogenic diabetes mellitus which also occurs secondary to pancreatic cancer [10]. The first category study of chronic pancreatitis (CP) in the United States is Prospective Evaluation of Chronic Pancreatitis for Epidemiologic and Translational Studies [27], and to better understand the pathophysiology and mechanisms, biorepositories are needed [8]. Research in pancreatic cancer, chronic pancreatitis (CP) and pancreatogenic diabetes mellitus were impeded due to inability in obtaining pancreatic tissue for study, genetic testing and biomarker development [23]. Genetic testing has been increasing clinical practice for the identification of susceptible genes for diseases [11]. The treatment of chronic pancreatitis includes pancreatectomy with autologous islet cell transplantation which is offered only in selected centres worldwide [22].

2.2

Causes

The factors responsible for chronic pancreatitis include smoking, autoimmune responses, genetics, alcohol, obstructive mechanisms, metabolic abnormalities and idiopathic mechanisms, whereas for acute pancreatitis, it includes autoimmune

2

Pancreatitis: Clinical Aspects of Inflammatory Phenotypes

29

responses, use of alcohol, hypertriglyceridaemia, biliary hypercalcaemia and hereditary/genetic abnormalities.

2.3

Acute Pancreatitis

Amongst gastrointestinal system diseases, one of the major forms is acute pancreatitis (AP) [28]. It is a serious disease possessing a guarded prognosis and aetiology which involves either the bladder or alcohol; it leads to trauma which is extremely rare and is often associated with intra-abdominal lesions [12]. Paediatric pancreatitis is understudied and an emerging field with an increasing incidence of disease and therefore extrapolated from adults [1]. A patient suffering from shortness of breath and acute pulmonary oedema upon admission at third day of hospitalisation was diagnosed with takotsubo cardiomyopathy [15]. Acute myocardial infarction incidence in acute pancreatitis is rare, and recognition and diagnosis of such may be clinically challenging [20]. Autoimmune systemic disease of IgG4-related disease defined as infiltration of IgG4-positive plasma cells in the affected organs leads to tissue inflammation which is rare and involves every organ [5]. Type I autoimmune pancreatitis is an IgG4-related systemic disease which exhibits as a pancreatic disorder, whereas type II is characterised by absence of IgG4-positive cells [3]. Toll like receptors initiate the excessive inflammatory reaction which makes the immune function unbalanced and also damages the function of many organs. Studies show TLR-4 is closely related to the occurrence and development of acute pancreatitis [5]. Acute pancreatitis is a necro-inflammatory disease diagnosed by elevated lipase levels and distinct imaging findings. One of the strategies to prevent recurrent attacks involves providing alcohol cessation counselling to patient [14]. Recent reports on fluid therapy, antibiotics, analgesics and the management of AP have led to improvements in clinical care [7]. Pancreatitis and Wilson’s disease were taken as examples in the study of ER stress in liver and pancreatic diseases [16]. Acute pancreatitis is caused by acinar cell injury, defective intracellular transport and duct obstruction. In acinar cell injury, agents such as alcohol, drugs, trauma and ischaemia cause activation of digestive enzymes through release of intracellular proenzymes and lysosomal hydrolases. In defective intracellular transport and metabolic injury, alcohol intake leads to acinar cell injury through delivery of proenzymes to lysosomal compartment and intracellular activation of enzymes. In duct obstruction and ampullary obstruction, chronic alcoholism, ductal concentration and intestinal oedema lead to impaired blood flow and ischemia and finally to acinar cell injury. This acinar cell injury leads to lesions such as intestinal inflammation, proteolysis, necrosis and haemorrhage (Fig. 2.1).

30

N. S. N. Chaitanya and A. BM Reddy

Fig. 2.1 Acute pancreatitis is mainly due to acinar cell injury which is caused by alcohol, smoking, metabolic injury and duct obstruction

2.4

Chronic Pancreatitis

Pancreatic inflammation with progressive loss of exocrine and endocrine gland compartments leading to atrophy or replacement with fibrous tissue in the affected area is known as chronic pancreatitis. Consequences of chronic pancreatitis include recurrent abdominal pain or constant abdominal pain, maldigestion and diabetes mellitus [13]. Characteristic features of chronic pancreatitis (CP) include pancreatic inflammation, scarring and fibrosis [17]. Clinical manifestations include pain in the abdominal region and insufficiency in exocrine and endocrine glands [18]. Even though acute pancreatitis and chronic pancreatitis were distinct [26], reports state that acute pancreatitis, recurrent acute pancreatitis and chronic pancreatitis represent a disease continuum [2, 21]. The risk factors include idiopathic, hereditary, autoimmune, severe, acute, pancreatitis-associated, chronic and obstructive pancreatitis [6]. This classification was based on individual’s risk of developing chronic pancreatitis [4] (Fig. 2.2).

2

Pancreatitis: Clinical Aspects of Inflammatory Phenotypes

31

Fig. 2.2 Chronic pancreatitis is caused by factors such as idiopathic, hereditary, autoimmune, severe, acute and obstructive pancreatitis

2.5

Diagnosis

Primary onset of acute pancreatitis was not detected till symptoms appear, while later stages were detected by fibrosis, distortion of the pancreatic ducts, pseudocyst and portal vein thrombosis or splenic vein thrombosis [13]. Patients with a long-standing history of alcohol or tobacco use are at risk for chronic pancreatitis which helps in diagnosis, and management involves reduced usage of alcohol and tobacco, modifications in diet and treatment of pancreatic insufficiency. Pancreatic neoplasms of acute, chronic and autoimmune pancreatitis were detected using diffusionweighted imaging. Ductal anomalies and characterisation of pancreatic cystic lesions were evaluated in acute and chronic pancreatitis using secretin-enhanced MR cholangiography [25]. Volatile organic compounds (VOC) were detected using field asymmetric waveform ion mobility spectrometry (FAIMS) by comparing healthy controls with pancreatic cancer from a urine sample [19]. RDW was not specific, but the CRP/albumin ratio is an inexpensive and reliable marker in prognosis of AP [28]. Amongst coronary imaging techniques, optimal cohesion tomography helps us

32

N. S. N. Chaitanya and A. BM Reddy

to observe coronary thrombus, inflammatory processes and presence of acute myocardial infarction. Intravascular imaging delineates details of intracoronary pathology which cannot be clearly predicted on standard fluoroscopy [20].

2.6

Treatment

Treatment includes medical, radiological, endoscopic and surgical [13]. Fluid therapy is used for acute pancreatitis. Endotherapy serves to achieve symptom remission and also reduces rates of recurrence, abdominal pain, hospitalisation and surgical arbitration [9].

2.7

Conclusion

Management of acute pancreatitis by fluid therapy yields better outcome. Acute pancreatitis complications can be prevented with limited usage of antibiotics as prophylaxis which shows beneficial effects. Severe acute pancreatitis benefits from ERCP, while mild patients are recommended to undergo single-stage laparoscopic cholecystectomy as they get benefit from single-stage and bile duct exploration which reduces length of hospitalisation and need for recurrent admissions. Acute pancreatitis is intent by alcohol, smoking and duct obstruction which proceeds to chronic pancreatitis which is caused by alcohol, obstruction of the pancreatic duct and hyperlipidemia. Progressive understanding of this disease had enhanced and changed the approach to management. Management of chronic pancreatitis involves counselling the patient regarding alcohol and tobacco and addressing the problems of malnutrition and osteoporosis. Acknowledgements NSNC acknowledges the ICMR for SRF fellowship and Laboratory of Cell Signalling supported partially by the University of Hyderabad through institutional funding from DST-FIST-II and PURSE, and ABMR acknowledges funding from DBT-RNAi, DAE-BRNS and DST-SERB, Government of India, through extramural research grants.

References 1. Abu-El-Haija M, Lowe ME (2018) Pediatric pancreatitis-molecular mechanisms and management. Gastroenterol Clin N Am 47(4):741–753 2. Ahmed Ali U et al (2016) Risk of recurrent pancreatitis and progression to chronic pancreatitis after a first episode of acute pancreatitis. Clin Gastroenterol Hepatol 14(5):738–746 3. Akshintala VS, Singh VK (2019) Management of Autoimmune Pancreatitis. Clin Gastroenterol Hepatol 4. Conwell DL et al (2014) American pancreatic association practice guidelines in chronic pancreatitis: evidence-based report on diagnostic guidelines. Pancreas 43(8):1143–1162 5. Damas F et al (2019) IgG4-related pleural disease in a patient with a history of unknown origin acute pancreatitis: a case report and review of the literature. Acta Clin Belg:1–4

2

Pancreatitis: Clinical Aspects of Inflammatory Phenotypes

33

6. Etemad B, Whitcomb DC (2001) Chronic pancreatitis: diagnosis, classification, and new genetic developments. Gastroenterology 120(3):682–707 7. Faghih M et al (2019) New advances in the treatment of acute pancreatitis. Curr Treat Options Gastroenterol 17(1):146–160 8. Fisher WE et al (2018) Standard operating procedures for biospecimen collection, processing, and storage: from the consortium for the study of chronic pancreatitis, diabetes, and pancreatic Cancer. Pancreas 47(10):1213–1221 9. Guo A, Poneros JM (2018) The role of Endotherapy in recurrent acute pancreatitis. Gastrointest Endosc Clin N Am 28(4):455–476 10. Hart PA et al (2018) Evaluation of a mixed meal test for diagnosis and characterization of PancrEaTogEniC DiabeTes secondary to pancreatic Cancer and chronic pancreatitis: rationale and methodology for the DETECT study from the consortium for the study of chronic pancreatitis, diabetes, and pancreatic Cancer. Pancreas 47(10):1239–1243 11. Hohmann M et al (2018) Practical genetic testing in gastroenterology. Dtsch Med Wochenschr 143(20):1477–1480 12. Khaoula Y et al (2018) Blunt abdominal trauma causing acute pancreatitis: presentation of the case study. Pan Afr Med J 30:126 13. Kleeff J et al (2017) Chronic pancreatitis. Nat Rev Dis Primers 3:17060 14. Klochkov A, Sun Y (2019) Alcoholic Pancreatitis. StatPearls. Treasure Island (FL) 15. Koop AH et al (2018) Acute pancreatitis-induced takotsubo cardiomyopathy and cardiogenic shock treated with a percutaneous left ventricular assist device. BMJ Case Rep 2018 16. Lukas J et al (2019) Role of endoplasmic reticulum stress and protein misfolding in disorders of the liver and pancreas. Adv Med Sci 64(2):315–323 17. Majumder S, Chari ST (2016) Chronic pancreatitis. Lancet 387(10031):1957–1966 18. Muniraj T et al (2014) Chronic pancreatitis, a comprehensive review and update. Part I: epidemiology, etiology, risk factors, genetics, pathophysiology, and clinical features. Dis Mon 60(12):530–550 19. Nissinen SI et al (2019) Detection of pancreatic Cancer by urine volatile organic compound analysis. Anticancer Res 39(1):73–79 20. Sanghvi S et al (2019) Coronary thrombosis in acute pancreatitis. J Thromb Thrombolysis 47 (1):157–161 21. Sankaran SJ et al (2015) Frequency of progression from acute to chronic pancreatitis and risk factors: a meta-analysis. Gastroenterology 149(6):1490–1500 e1491 22. Schrope B (2018) Total pancreatectomy with autologous islet cell transplantation. Gastrointest Endosc Clin N Am 28(4):605–618 23. Serrano J et al (2018) Consortium for the study of chronic pancreatitis, diabetes, and pancreatic Cancer: from concept to reality. Pancreas 47(10):1208–1212 24. Shah AP et al (2018) Acute pancreatitis: current perspectives on diagnosis and management. J Inflamm Res 11:77–85 25. Siddiqui N et al (2018) Advanced MR imaging techniques for pancreas imaging. Magn Reson Imaging Clin N Am 26(3):323–344 26. Singer MV et al (1985) 2d symposium on the classification of pancreatitis. Marseilles, 28–30 March 1984. Acta Gastroenterol Belg 48(6):579–582 27. Yadav D et al (2018) PROspective evaluation of chronic pancreatitis for EpidEmiologic and translational StuDies: rationale and study design for PROCEED from the consortium for the study of chronic pancreatitis, diabetes, and pancreatic Cancer. Pancreas 47(10):1229–1238 28. Yilmaz EM, Kandemir A (2018) Significance of red blood cell distribution width and C-reactive protein/albumin levels in predicting prognosis of acute pancreatitis. Ulus Travma Acil Cerrahi Derg 24(6):528–531

3

Diabetes and Pancreatic Cancer: A Bidirectional Relationship Perspective Manoj Kumar Gupta, Vemula Sarojamma, and Ramakrishna Vadde

Abstract

An attempt was made to understand the bidirectional causal relationship between diabetes mellitus and pancreatic cancer. One theory claims that diabetes mellitus is a risk factor for developing pancreatic cancer. Supporting this, several studies have reported that four type 2 diabetes-associated polymorphisms, namely, rs8050136 of FTO, rs1387153 of MTNR1B, rs11039149 of MADD and rs780094 of GCKR, are associated risk factors for pancreatic cancer. Enhanced concentration of circulating C-peptide/insulin during hyperglycaemia is also reported to be the risk factors for developing both colorectal and pancreatic cancer. However, another theory claims that pancreatic cancer is risk factor for developing diabetes mellitus. Insulin sensitivity in pancreatic cancer patients is reported to increase after tumour resection supporting pancreatic cancer as a causal agent of DM. Incidence of type 2 diabetes (T2DM) leading to pancreatic cancer is more in comparison to type 1 diabetes. Successful treatment of obesity and T2DM decreases pancreatic cancer risk. Nevertheless, treatment of T2DM with insulin, insulin analogues, and insulin secretagogues enhances the risk for pancreatic cancer. Metformin, an antidiabetic and antineoplastic drug, is reported to reduce the risk of pancreatic cancer and should be considered as first-line therapy in all new-onset diabetes patients. Apart from metformin, antimalarial drug, namely, chloroquine, and other nutritional supplements, like limonene and melatonin, are also utilised in pancreatic cancer prevention and treatment. In near future, this information may be utilised in reversing the increasing incidence of pancreatic cancer. M. K. Gupta · R. Vadde (*) Department of Biotechnology & Bioinformatics, Yogi Vemana University, Kadapa, Andhra Pradesh, India e-mail: [email protected] V. Sarojamma Department of Microbiology, Sri Venkateswara Medical College, Tirupathi, Andhra Pradesh, India # Springer Nature Singapore Pte Ltd. 2019 G. P. Nagaraju, A. BM Reddy (eds.), Exploring Pancreatic Metabolism and Malignancy, https://doi.org/10.1007/978-981-32-9393-9_3

35

36

M. K. Gupta et al.

Keyword

Diabetes · Pancreatic cancer · Metformin · Insulin resistance · Hyperglycaemia

3.1

Introduction

Diabetes mellitus (DM) is a complex metabolic disorder clinically described by high blood glucose concentration because of reduced insulin production, insulin resistance, or both [89]. Though more than 50 forms of DM identified by the American Diabetes Association, DM is categorised mainly into two types, namely, type 1 diabetes mellitus (T1DM) and type 2 diabetes mellitus (T2DM) [8]. T1DM is an autoimmune destruction of pancreatic β-cells. Unlike T1DM, T2DM has incapability of pancreatic β-cells to produce sufficient insulin required by the body. T2DM is a heterogeneous disorder caused by both genetic and environmental factors including oxidative stress [89, 107, 108]. T2DM constitutes about 90–95% of all forms of DM in human. In general population, incidence of T2DM increases with age [89]. Earlier studies have reported that chronic hyperglycaemia lead to various forms of cancer, for instance, urinary tract, liver and kidney cancer [2, 56, 145]. However, incidence of T2DM leading to cancer is more in comparison to T1DM [142]. Indeed, only few studies have reported association between T1DM and cancer [47]. Studies have also reported that women with DM are more prone to breast cancer in comparison to the women without DM [16, 76]. Relationship between DM and some cancer may be because they share same risk factor; for instance, obesity and aging are causing agent for both DM and pancreatic cancer [47]. Growing BMI increases the risk ratio for cancer. Obese patients having BMI greater than 30 kg/m2 are more prone to cancers in comparison to overweight (BMI between 25 kg/m2 and 30 kg/m2) patients [19]. Apart from insulin resistance, obesity also enhances oestrogen production, which in turn may increase the risk of oestrogen-dependent tumours. Weight gain is often associated with high risk for breast, cervix and endometrium cancer in women [37, 75, 131]. Recently, our team reported four T2D genes, namely, TCF7L2, CCL2, ELMO1 and VEGFA, along with FOS which plays an important role in causing T2D and its associated disorders, like rheumatoid arthritis, nephropathy, neuropathy and cancer via p53 or Wnt signalling pathways [54]. Thus, both DM and cancer are complex processes, and normal cell must undergo several transformations before turning into cancerous cell. For example, activation of KRAS and suppression of CDKN2A and SMAD4 initiate pancreatic cancer progression [34, 109]. Most plausible molecular functions by which DM influences neoplastic process are chronic inflammation, hyperglycaemia and hyperinsulinaemia [47]. A study reported that diabetes-associated diseases like steatosis, cirrhosis and nonalcoholic fatty liver disease are responsible for inducing liver cancer [47]. About 80% of pancreatic cancer is reported to occur because of glucose intolerance [79]. Incidence of DM amongst pancreatic cancer patients is high in comparison to other cancers [119]. About 50–80% of pancreatic cancer patients are reported

3

Diabetes and Pancreatic Cancer: A Bidirectional Relationship Perspective

37

to be diagnosed with DM at least 3 years prior to pancreatic cancer diagnosis [98, 100, 135]. A distinct segment on pancreatic cancer in the American Cancer Society’s ‘Cancer Facts & Figures 2013’ [123] suggests, ‘About 25% of patients with pancreatic cancer have diabetes mellitus at diagnosis, and roughly another 40% have pre-diabetes’, that ‘patients with long-term (five or more years) type II diabetes have a 50% increased risk of pancreatic cancer’ and ‘pancreatic cancer can cause diabetes, and sometimes diabetes is an early sign of the tumour’ proposing bidirectional relationship between DM and pancreatic cancer [111]. Overall 5 year survival rate of pancreatic cancer patients is 7–8% and is reported to be the fourth leading cause of cancer-related deaths [67, 106, 121]. As the causal mechanism of this bidirectional relationship between DM and pancreatic cancer is still a topic of debate, in this chapter, authors discussed about the bidirectional relationship between DM and pancreatic cancer. In the beginning, we have discussed about possible pathways, namely, insulin resistance, hyperglycaemia and chronic inflammation, by which both are interlinked. Later, we have discussed about various evidences supporting bidirectional relationship between them. Finally, we have discussed about various drugs or nutrient supplement that can be utilised in the treatment of pancreatic cancer. In near future, these information may be utilised in treatment of both DM and pancreatic cancer.

3.2

Plausible Link Between DM and Cancer

Insulin resistance, hyperglycaemia and chronic inflammation are three important mechanisms that interlink DM with cancer [138]. During insulin resistance, which is a common feature of T2DM, insulin level in circulating blood frequently increases, which in turn causes hyperinsulinaemia [138]. Hyperinsulinaemia may initiate development of cancer either by binding insulin to insulin receptor (IR) or increasing the concentration of circulating insulin growth factor (IGF)-1 [112]. Insulin signal transduction is regulated via isoforms of IR, namely, IR-A and IR-B. IR-A has affinity for both insulin and insulin growth factors (IGFs) and is expressed mainly in cancer cells and foetal tissues [138]. IR-B is insulin specific and is mainly associated with glucose homeostasis. When insulin ligates with IR-A, it initiates pro-growth mitogenic effect which in turn initiates carcinogenesis. Hyperinsulinaemia also enhances the expression of IGF-1 in the liver, further activating IGF-1 receptor, which in turn stimulates cell proliferation causing cancer [33, 94]. Hyperglycaemia may induce cancer either by ‘direct effect’ or ‘indirect effect’ [138]. During ‘direct effect’, hyperglycaemia directly affects tumour cells by inducing mutations, enhancing invasion as well as migration and further renewing cancerrelated signalling pathways. During ‘indirect effect’, initial action takes place at another organs/tissue, which will later influence tumour cells via stimulating production of circulating growth factors (insulin/IGF-1) as well as inflammatory cytokines [138]. Hyperglycaemia is reported to enhance Wnt/β-catenin signalling, which is a key cancer-associated pathway, by permitting nuclear retention as well as aggregation of transcriptionally active β-catenin [6, 26].

38

M. K. Gupta et al.

Chronic inflammation and chronic oxidation are reported to work together. Hyperglycaemia produced reactive oxygen species (ROS) which damages protein, lipids and DNA, which in turn leads to cancer [138]. ROS-induced chronic inflammation enhances TNF-α concentration, which in turn stimulates nuclear factorkappa B (NF-κB), followed by development as well as progression of numerous tumours [31, 60, 69]. Thus, continuous exposure to oxidative stress as well as chronic inflammation transforms susceptible cells to cancerous cell [133].

3.3

Risk Factors for DM

As stated above, T2DM is a complex disorder that develops due to interaction between strong hereditary component and environmental factors. Though the environmental factors play significant role in the development of diabetes, their impact varies in each individual [5]. Even under the same environmental conditions, some people are more prone to T2DM, while some are not, suggesting hereditary factor plays important role in the development of T2DM. The advancement of genetic technologies had enabled researcher to identify various T2DM loci via linkage studies or candidate gene approach [5]. Linkage studies identified two important T2DM genes, namely, calpain-10 gene (CAPN10) [58, 61] and TCF7L2 [35, 51]. Few important T2DM genes identified via candidate gene approach are insulin receptor substrate 1 (IRS1) and IRS2, peroxisome proliferator-activated receptor gamma (PPARG), potassium inwardly rectifying channel, subfamily J, member 11 (KCNJ11), HNF1 homeobox B (HNF1B), HNF4A and Wolfram syndrome 1 (wolframin) [5]. Recent genome-wide association study (GWAS) in Asian population by Hu and team reported that SNPs in PPARG (rs1801282), IGF2BP2 (rs7651090), KCNJ11 (rs5219), CDKN2A-CDKN2B (rs564398 and rs10811161), CDKAL1 (rs10946398, rs7754840, rs9460546, rs7756992 and rs9465871), SLC30A8 (rs13266634) and IDE-KIF11-HHEX (rs10509645, rs1111875 and rs10748582) are risk factor for T2DM [146]. Another GWAS analysis reported rs5219 polymorphism of the KCNJ11 gene is a risk factor for developing T2DM in Caucasians [116], European and East Asia, but not in Mongolian [1, 24, 95], Indian [23, 117] or Ashkenasi Jewish [92]. The KCNJ11 gene, a member of the potassium channel gene family, is situated at 11p15.1. Potassium channels are found in most mammalian cells and are associated with many physiological activities including apoptosis, immunomodulation and insulin secretion. KCNJ11 encodes inward-rectifier potassium ion channel (Kir6.2). Kir6.2 binds with sulfonylurea receptor 1 (SUR1) to form KATP channel, which regulates production and secretion of insulin via glucose metabolism. Mutation in KCNJ11 (rs5219) gene disrupts the normal function of KATP channel and thus production and secretion of insulin which consequently causes T2DM in human [49, 129, 130]. Earlier studies have reported that pregnant women having T allele are at higher risk of developing gestational diabetes mellitus (GDM) ([87, 78]. Thus, the rs5219 (p.Lys23Glu) variation in the 11p15.1 region plays vital role in causing T2DM, thus making it a popular marker for controlling T2DM.

3

Diabetes and Pancreatic Cancer: A Bidirectional Relationship Perspective

39

GWAS analysis in Asian [25, 27, 115] and European [27] human populations reported that rs1535500 polymorphism of the KCNK16 gene is a risk factor for developing T2DM. KCNK16 encodes the potassium two-pore domain channel subfamily K member 16 (alias TALK-1) [91] and is situated at 6p21.2. These channels are mainly expressed in human and mouse pancreatic β-cells and generate outwardly rectifying, non-inactivating K+ flows, which increases with increase in extracellular pH [48]. This polymorphism increases TALK-1 channel activity, disturbing normal membrane potential of β-cell, Ca2+ influx and insulin secretion which ultimately leads to T2DM [132]. Recently, through computational approach, we have reported that polymorphism rs13266634 in ZnT8 transporter disrupts the normal movement and accumulation of zinc into intracellular vesicles from cytoplasm, which in turn have negative impact on the production, storage and secretion of insulin by β-cells of the pancreas and thus causes T2D in human [55]. Reactive oxygen species, like singlet oxygen, are also reported to increase the activity of TALK-1 channel [36, 48, 57, 68]. Vierra and team suggested that inhibition of β-cell TALK-1 channels may be a novel therapeutic approach for reducing hyperglycaemia in T2DM [132]. Recently, Someya and team reported that Japanese college students having BMI > 22.0 kg/m2 are more prone to diabetes later in life [124]. Adela and team via nonparametric-based machine learning approach identified numerous cellular network proteins, like IL1R2, PTPN1, IRS2, AKT1, INSR, LEPR, IRS1, IL6R, MYD88 and PCSK9 which are associated with insulin resistance, atherosclerosis and inflammation [3]. Epicardial fat volume in early-onset T2DM is signature of reduced renal function in future [110]. Zhang and team reported that obstructive sleep apnoea disorder has negative influence on glucose metabolism and causes T2DM [144].

3.4

Risk Factors for Pancreatic Cancer

Carcinogens are the most important factors for inducing pancreatic cancer. Smoking tobacco accounts for about 20 to 30% of pancreatic cancer [82]. Exposure to environmental tobacco smoke in childhood days enhances the risk of pancreatic cancer [134]. Analysis of Nurses’ Health Study data proposes that in utero exposure to smoking is also a plausible cause for pancreatic cancer [9]. Exposure to chlorinated hydrocarbon as well as organochloride compounds also enhances the risk of pancreatic cancer [97, 103]. Infection via Helicobacter pylori [105] and hepatitis B [59] and tropical pancreatitis, chronic pancreatitis and hereditary pancreatitis are also reported to be risk factors for pancreatic cancer [71, 83]. People belonging to older age and with Ashkenazi Jewish and African-American ancestry are more prone to pancreatic cancer [82]. Incidence is lowest amongst African, Southeast Asia and Indian populations [67]. Familial history also plays significant role in the development of 10–20% of pancreatic carcinomas [82]. Mutation in BRCA2 [17], PALB2 [17], p16 [50], STK11/ LKB [46], PRSS1 [83], SPINK1 [83], IGF-1 [127] and ATM [113] also reported to be associated with pancreatic cancer. Recently, one study observed decreased levels of

40

M. K. Gupta et al.

HbA1c in benign pancreatic cancer patients in comparison with malignant patients, and recommending HbA1c may also serve as a biomarker for pancreatic cancer [38]. Chronic pancreatitis is also reported to increase the risk of pancreatic cancer [67]. DM patients are also reported to have 30% risk of developing pancreatic cancer [67]. Association between DM and pancreatic cancer has been known for almost 200 years. In 1833, Bright reported that DM patients died of pancreatic cancer within 6 month [18]. In late 1950s, Green and team reported that almost 50% of pancreatic cancer patients at the Mayo clinic have prior history of glycosuria or DM [52]. Positive relationship between pancreatic cancer and specifically high body mass index (BMI), obesity and centralised fat distribution has also been reported. Insulin secretagogues, for instance, sulfonylurea that stimulated β-cells to release insulin, are also reported to induce various cancers in T2DM patients [29, 86]. Compensatory hyperinsulinaemia, insulin resistance and high levels of circulating insulin-like growth factors (IGF) are the most important factors underlying in the association between T2DM and pancreatic cancer [79]. Animal studies have reported that initiation of pancreatic cancer is dependent on the turnover of islet cells. For instance, in hamsters, initiation of islet cell proliferation increases carcinogenesis of pancreatic duct [38]. Though earlier studies have reported that 50% of pancreatic cancer is reported to have prior DM history, it still remains unclear if early onset of pancreatic cancer initiates DM or DM is a risk factor for pancreatic cancer.

3.5

Evidences Proposing DM Is a Risk Factor for Pancreatic Cancer

Everhat and team reanalysed 20 epidemiology cases that have studied relation between DM and pancreatic cancer via meta-analysis and reported that pancreatic cancer may develop due to complication in DM [39]. Another meta-analysis of studies of breast, colon and rectum, endometrium and pancreatic cancers suggested that increased concentration of circulating C-peptide/insulin and other markers of hyperglycaemia is the risk factors for developing colorectal and pancreatic cancer [102]. One population-based study in the United States with 2153 controls and 526 pancreatic cancer cases reveals significant positive trend in risk (P ¼ 0.016) with increasing years before diagnosis of cancer [122]. In T2DM patients, as pancreas is exposed to substantial hyperinsulinaemia for several years, researchers believe that chronic hyperinsulinaemia may stimulate the growth of pancreatic cancer [45]. Binding studies have observed insulin receptors on pancreatic cancer cell [40, 42, 43]. In vitro studies have reported insulin stimulates growth of hamster [43], rat [88] and human [12, 32, 74, 128, 136] pancreatic cancer cell lines. Apart from hyperinsulinaemia, increased concentration of free fatty acids is also reported to initiate pancreatic cancer [41]. Numerous GWAS have reported various genetic variants that are responsible for causing both DM and pancreatic cancer. Pierce and team have studied T2DM 37 risk alleles and reported that three polymorphisms, namely, rs8050136 of FTO, rs1387153 of MTNR1B and rs11039149 of MADD, are

3

Diabetes and Pancreatic Cancer: A Bidirectional Relationship Perspective

41

associated with high risk of pancreatic cancer [101]. Prizment and team reported that rs780094 polymorphisms of glucokinase regulator (GCKR), which is responsible for causing T2DM, may also be a risk factor for pancreatic cancer [104]. Steven and team reported that risk for T1DM and ‘young-onset’ DM patients to develop pancreatic cancer are two-time higher than non-DM individual [125]. Another form of DM, namely, pancreatogenic, or type 3c diabetes (T3cDM), constitutes 5–10% of Western DM populations and is responsible for causing several numerous other diseases. T3cDM is characterised by hepatic insulin resistance and is caused by deficits of both pancreatic polypeptide and insulin. Chronic pancreatitis, which is one of the most important risk factors of pancreatic cancer, is responsible for causing 70% T3cDM [28]. T3cDM is also responsible for causing development of metabolic bone disease and nutritional deficiencies [28]. However, contrasting results proposing DM is not a risk factor for pancreatic cancer have also been proposed by several researchers [20, 44, 53]. Human pancreatic cancer cell line SOJ-6 [128] and PANC-1 [12] are reported not to be affected via insulin.

3.6

Evidences Proposing Pancreatic Cancer Is a Risk Factor for DM

Most of the pancreatic cancer-related DM is diagnosed either simultaneously with pancreatic cancer or 2 years prior to diagnosis of pancreatic cancer suggesting recent onset of DM may be an early sign of pancreatic cancer [53, 118]. Numerous studies have reported peripheral insulin resistance in pancreatic cancer patients with DM [100, 118]. Though few, insulin resistance is also reported in non-DM pancreatic cancer patients with lesser degree [99]. Permert and team reported that insulin sensitivity in pancreatic cancer patients increases after tumour resection supporting pancreatic cancer as a causal agent of DM [99]. Basso and team identified 2030 MW diabetogenic peptide in serum of pancreatic cancer patients [11]. A number of investigators have studied insulin resistance at the organ, tissue and cellular levels in pancreatic cancer [10, 63, 80, 81]. In human skeletal muscles, during initial steps of the insulin signalling cascade, no significant difference amongst healthy controls and pancreatic cancer patients were reported in insulin receptor binding, tyrosine kinase activity and insulin receptor substrate-1 content [81]. However, downstream steps of the insulin signalling cascade, namely, phosphatidylinositol 3-kinase (PI3-K) activity and glucose transport, were reported to be impaired in pancreatic cancer patients [63]. Moreover, glycogen synthase activity is also reported to be reduced in skeletal muscle of both pancreatic cancer human and rodent suggesting insulin signalling cascade gets impaired at various stages of pancreatic cancer [80, 81, 99]. Islet dysfunction is another interesting feature of DM related with pancreatic cancer [137]. As the amount of islet tissues depleted via tumour is very small, islet dysfunction is less likely to decrease islet volume. Indeed, endocrine pancreatic function can be modulated even with large depletion of pancreatic tissue

42

M. K. Gupta et al.

[15]. Reduced insulin secretion in pancreatic cancer patient is due to reduce response to stimuli [21, 118]. Chemically induced pancreatic cancer in hamsters is also reported to impair release of normal glucose-stimulated insulin [4]. Ishikawa and team observed enhanced level of proinsulin, in comparison to insulin, in patients with pancreatic cancer patients, proposing that proinsulin maturation may also be affected by the tumour [64].

3.7

Drugs Utilized in the Treatment of Pancreatic Cancer

Thrombosis is a common problem in pancreatic cancer, and chronic pancreatic cancer causes venous thromboembolism [120, 139]. Anticoagulant treatment via aspirin is reported to improvise survival rate of pancreatic cancer patient via reducing thromboembolic complexity [84, 90, 114]. Two independent meta-analyses reported that metformin consumption could reduce the incident of pancreatic cancer by 46% [38, 143]. Metformin reduces blood glucose levels mainly via supressing hepatic glucose production, specifically hepatic gluconeogenesis, and increasing peripheral tissue insulin sensitivity [72]. Metformin is also reported to have a cardioprotective property, which in turn modulates fat metabolism as well as adipose tissue hormone production, mainly leptin [62]. Treatment of T2DM patients with metformin improvises endothelial functions, including soluble vascular cell adhesion molecule-1, tissue-type plasminogen activator and inhibitor, soluble E-selectin and vascular endothelial growth factor [30, 65, 85]. Majority of these molecules play significant role in angiogenesis, fibrosis and thrombosis. However, the important molecular mechanism of metformin’s effect on metabolism as well as growth inhibition is modulated via liver kinase B1 (LKB1)/50 AMPK (AMP-activated protein kinase) signalling pathway. In dormant cell, majority of the energy (in ATP form) is generated in mitochondria through oxidative phosphorylation. Metformin is reported to enhance glucose uptake as well as glycolysis and decreases mitochondrial ATP production, which in turn activates LKB1/AMPK signalling pathway further restricting the growth of pancreatic cancerous cell. Depletion of islet cells through streptozotocin, biguanide metformin and alloxan treatment restricts the induction of pancreatic cancer [13, 22, 96]. Chloroquine, antimalarial drug, is also reported to restrict the growth of pancreatic tumours via inhibiting ‘autophagy’ [140]. However, precaution must be taken with alcoholic pancreatic cancer patients. Nutritional supplements, like limonene [70]; melatonin [77]; vitamins A, C and E [14]; selenium [7]; vitamin K [93], vitamin B6 [126] and green tea [66] and some synthetic drugs, like polysaccharide K [73] and Ukrain (NSC-631570) [141], are also utilised in pancreatic cancer prevention and treatment. Hence, metformin, chloroquine and other nutritional supplements, like limonene and melatonin, may be utilised in pancreatic cancer prevention and treatment.

3

Diabetes and Pancreatic Cancer: A Bidirectional Relationship Perspective

3.8

43

Conclusion and Perspective

Earlier studies reported link between chronic DM and pancreatic cancer. However, exact mechanism linking DM and pancreatic cancer is a topic of debate. Some studies hypothesize that DM is a risk factor for pancreatic cancer, while other proposed pancreatic cancer is a risk factor for DM. Thus, in this review, an attempt was made to understand the bidirectional relationship between DM and pancreatic cancer. Insulin resistance, hyperglycaemia and chronic inflammation are three important factors that interlink DM with cancer. However, T2DM patients are more prone to pancreatic than T1DM patients. Successful treatment of obesity and T2DM and obesity decreases pancreatic cancer risk. Nevertheless, treatment of T2DM through with insulin analogues, and insulin secretagogues enhances the risk for pancreatic cancer. Metformin, an antidiabetic and antineoplastic drug, is reported to reduce the risk of pancreatic cancer and should be considered as first-line therapy in all new-onset DM patients over the age of 50. Apart from metformin, antimalarial drug, namely, chloroquine, and nutritional supplements, like limonene [70] and melatonin [77], are also utilised in pancreatic cancer prevention and treatment. In near future, these information may be utilised in treatment of both DM and pancreatic cancer as well as reversing the increasing incidence of pancreatic cancer. Besides, it seems that many aspects of molecular mechanism modulating bidirectional relationship between DM and pancreatic cancer are still unknown. Identification of significant genes, molecular pathways and environmental factors, through computational as well as in vivo/in vitro studies, may provide us with an opportunity to reduce cancer at higher speed. Conflicts of Interest Authors declare no conflicts of interest.

References 1. Abdelhamid I, Lasram K, Meiloud G, Ben Halim N, Kefi R, Samb A et al (2014) E23K variant in KCNJ11 gene is associated with susceptibility to type 2 diabetes in the Mauritanian population. Prim Care Diabetes 8:171–175. https://doi.org/10.1016/j.pcd.2013.10.006 2. Adami H-O, McLaughlin J, Ekbom A, Berne C, Silverman D, Hacker D et al (1991) Cancer risk in patients with diabetes mellitus. Cancer Causes Control 2:307–314. https://doi.org/10. 1007/BF00051670 3. Adela R, Reddy PNC, Ghosh TS, Aggarwal S, Yadav AK, Das B et al (2019) Serum protein signature of coronary artery disease in type 2 diabetes mellitus. J Transl Med 17:17. https://doi. org/10.1186/s12967-018-1755-5 4. Ahrén B, Andrén-Sandberg A (1993) Glucose tolerance and insulin secretion in experimental pancreatic cancer in the Syrian hamster. Res Exp Med Z Gesamte Exp Med Einschl Exp Chir 193:21–26 5. Ali O (2013) Genetics of type 2 diabetes. World J Diabetes 4:114–123. https://doi.org/10. 4239/wjd.v4.i4.114 6. Anagnostou SH, Shepherd PR (2008) Glucose induces an autocrine activation of the Wnt/betacatenin pathway in macrophage cell lines. Biochem J 416:211–218. https://doi.org/10.1042/ BJ20081426

44

M. K. Gupta et al.

7. Armstrong T, Strommer L, Ruiz-Jasbon F, Shek FW, Harris SF, Permert J et al (2007) Pancreaticoduodenectomy for peri-ampullary neoplasia leads to specific micronutrient deficiencies. Pancreatol Off J Int Assoc Pancreatol IAP Al 7:37–44. https://doi.org/10.1159/ 000101876 8. Association AD (2014) Diagnosis and classification of diabetes mellitus. Diabetes Care 37: S81–S90. https://doi.org/10.2337/dc14-S081 9. Bao Y, Giovannucci E, Fuchs CS, Michaud DS (2009) Passive smoking and pancreatic cancer in women: a prospective cohort study. Cancer Epidemiol Biomark Prev Publ Am Assoc Cancer Res Cosponsored Am Soc Prev Oncol 18:2292–2296. https://doi.org/10.1158/10559965.EPI-09-0352 10. Basso D, Brigato L, Veronesi A, Panozzo MP, Amadori A, Plebani M (1995) The pancreatic cancer cell line MIA PaCa2 produces one or more factors able to induce hyperglycemia in SCID mice. Anticancer Res 15:2585–2588 11. Basso D, Valerio A, Seraglia R, Mazza S, Piva MG, Greco E et al (2002) Putative pancreatic cancer-associated diabetogenic factor: 2030 MW peptide. Pancreas 24:8–14 12. Beauchamp RD, Lyons RM, Yang EY, Coffey RJ, Moses HL (1990) Expression of and response to growth regulatory peptides by two human pancreatic carcinoma cell lines. Pancreas 5:369–380 13. Bergmann U, Funatomi H, Yokoyama M, Beger HG, Korc M (1995) Insulin-like growth factor I overexpression in human pancreatic cancer: evidence for autocrine and paracrine roles. Cancer Res 55:2007–2011 14. Bjelakovic G, Nikolova D, Simonetti RG, Gluud C (2004) Antioxidant supplements for prevention of gastrointestinal cancers: a systematic review and meta-analysis. Lancet Lond Engl 364:1219–1228. https://doi.org/10.1016/S0140-6736(04)17138-9 15. Bonner-Weir S, Trent DF, Weir GC (1983) Partial pancreatectomy in the rat and subsequent defect in glucose-induced insulin release. J Clin Invest 71:1544–1553 16. Bota M, Autier P, Boyle P (2018) The risk of breast cancer in women with diabetes. Diabetes 67:180-OR. https://doi.org/10.2337/db18-180-OR 17. Breast Cancer Linkage Consortium (1999) Cancer risks in BRCA2 mutation carriers. J Natl Cancer Inst 91:1310–1316 18. Bright R (1833) Cases and observations connected with disease of the pancreas and duodenum. Medico-Chir Trans 18:1–56 19. Calle EE, Kaaks R (2004) Overweight, obesity and cancer: epidemiological evidence and proposed mechanisms. Nat Rev Cancer 4:579–591. https://doi.org/10.1038/nrc1408 20. Calle EE, Murphy TK, Rodriguez C, Thun MJ, Heath CW (1998) Diabetes mellitus and pancreatic cancer mortality in a prospective cohort of United States adults. Cancer Causes Control CCC 9:403–410 21. Cersosimo E, Pisters PW, Pesola G, McDermott K, Bajorunas D, Brennan MF (1991) Insulin secretion and action in patients with pancreatic cancer. Cancer 67:486–493 22. Chang M-C, Chang Y-T, Su T-C, Yang W-S, Chen C-L, Tien Y-W et al (2007) Adiponectin as a potential differential marker to distinguish pancreatic cancer and chronic pancreatitis. Pancreas 35:16–21. https://doi.org/10.1097/MPA.0b013e3180547709 23. Chavali S, Mahajan A, Tabassum R, Dwivedi OP, Chauhan G, Ghosh S et al (2011) Association of variants in genes involved in pancreatic β-cell development and function with type 2 diabetes in north Indians. J Hum Genet 56:695–700. https://doi.org/10.1038/jhg.2011.83 24. Chen G, Xu Y, Lin Y, Lai X, Yao J, Huang B et al (2013) Association study of genetic variants of 17 diabetes-related genes/loci and cardiovascular risk and diabetic nephropathy in the Chinese she population. J Diabetes 5:136–145. https://doi.org/10.1111/1753-0407.12025 25. Cho YS, Chen C-H, Hu C, Long J, Ong RTH, Sim X et al (2011) Meta-analysis of genomewide association studies identifies eight new loci for type 2 diabetes in East Asians. Nat Genet 44:67–72. https://doi.org/10.1038/ng.1019

3

Diabetes and Pancreatic Cancer: A Bidirectional Relationship Perspective

45

26. Chocarro-Calvo A, García-Martínez JM, Ardila-González S, De la Vieja A, García-Jiménez C (2013) Glucose-induced β-catenin acetylation enhances Wnt signaling in cancer. Mol Cell 49:474–486. https://doi.org/10.1016/j.molcel.2012.11.022 27. Consortium DiaGRAM (DIAGRAM), Consortium AGENT 2 D (AGEN-T, Consortium SAT 2 D (SAT2D), Consortium MAT 2 D (MAT2D), Consortium T 2 DGE by N sequencing in multi-ES (T2D-G), Mahajan A, et al. (2014) Genome-wide trans-ancestry meta-analysis provides insight into the genetic architecture of type 2 diabetes susceptibility. Nat Genet 46:234–244. https://doi.org/10.1038/ng.2897 28. Cui Y, Andersen DK (2011) Pancreatogenic diabetes: special considerations for management. Pancreatology 11:279–294. https://doi.org/10.1159/000329188 29. Currie CJ, Poole CD, EAM G (2009) The influence of glucose-lowering therapies on cancer risk in type 2 diabetes. Diabetologia 52:1766–1777. https://doi.org/10.1007/s00125-0091440-6 30. De Jager J, Kooy A, Lehert P, Bets D, Wulffelé MG, Teerlink T et al (2005) Effects of shortterm treatment with metformin on markers of endothelial function and inflammatory activity in type 2 diabetes mellitus: a randomized, placebo-controlled trial. J Intern Med 257:100–109. https://doi.org/10.1111/j.1365-2796.2004.01420.x 31. DiDonato JA, Mercurio F, Karin M (2012) NF-κB and the link between inflammation and cancer. Immunol Rev 246:379–400. https://doi.org/10.1111/j.1600-065X.2012.01099.x 32. Ding XZ, Fehsenfeld DM, Murphy LO, Permert J, Adrian TE (2000) Physiological concentrations of insulin augment pancreatic cancer cell proliferation and glucose utilization by activating MAP kinase, PI3 kinase and enhancing GLUT-1 expression. Pancreas 21:310–320 33. Djiogue S, Nwabo Kamdje AH, Vecchio L, Kipanyula MJ, Farahna M, Aldebasi Y et al (2013) Insulin resistance and cancer: the role of insulin and IGFs. Endocr Relat Cancer 20:1–17. https://doi.org/10.1530/ERC-12-0324 34. Doucas H, Garcea G, Neal CP, Manson MM, Berry DP (2006) Chemoprevention of pancreatic cancer: a review of the molecular pathways involved, and evidence for the potential for chemoprevention. Pancreatol Off J Int Assoc Pancreatol IAP Al 6:429–439. https://doi.org/ 10.1159/000094560 35. Duggirala R, Blangero J, Almasy L, Dyer TD, Williams KL, Leach RJ et al (1999) Linkage of type 2 diabetes mellitus and of age at onset to a genetic location on chromosome 10q in Mexican Americans. Am J Hum Genet 64:1127–1140. https://doi.org/10.1086/302316 36. Duprat F, Girard C, Jarretou G, Lazdunski M (2005) Pancreatic two P domain K+ channels TALK-1 and TALK-2 are activated by nitric oxide and reactive oxygen species. J Physiol 562:235–244. https://doi.org/10.1113/jphysiol.2004.071266 37. Eliassen AH, Colditz GA, Rosner B, Willett WC, Hankinson SE (2006) Adult weight change and risk of postmenopausal breast Cancer. JAMA 296:193–201. https://doi.org/10.1001/jama. 296.2.193 38. El-Jurdi NH, Saif MW (2013) Diabetes and pancreatic cancer. JOP J Pancreas 14:363–366. https://doi.org/10.6092/1590-8577/1656 39. Everhart J, Wright D (1995) Diabetes mellitus as a risk factor for pancreatic cancer. A metaanalysis JAMA 273:1605–1609 40. Fisher WE, Boros LG, O’Dorisio TM, O’Dorisio MS, Schirmer WJ (1995a) GI hormonal changes in diabetes influence pancreatic cancer growth. J Surg Res 58:754–758. https://doi. org/10.1006/jsre.1995.1119 41. Fisher WE, Boros LG, Schirmer WJ (1995b) Reversal of enhanced pancreatic cancer growth in diabetes by insulin. Surgery 118:453–457. discussion 457-458 42. Fisher WE, Boros LG, Schirmer WJ (1996) Insulin promotes pancreatic cancer: evidence for endocrine influence on exocrine pancreatic tumors. J Surg Res 63:310–313. https://doi.org/10. 1006/jsre.1996.0266

46

M. K. Gupta et al.

43. Fisher WE, Muscarella P, Boros LG, Schirmer WJ (1998) Variable effect of streptozotocindiabetes on the growth of hamster pancreatic cancer (H2T) in the Syrian hamster and nude mouse. Surgery 123:315–320 44. Frye JN, Inder WJ, Dobbs BR, Frizelle FA (2000) Pancreatic cancer and diabetes: is there a relationship? A case-controlled study. Aust N Z J Surg 70:722–724 45. Gapstur SM, Gann PH, Lowe W, Liu K, Colangelo L, Dyer A (2000) Abnormal glucose metabolism and pancreatic cancer mortality. JAMA 283:2552–2558 46. Giardiello FM, Brensinger JD, Tersmette AC, Goodman SN, Petersen GM, Booker SV et al (2000) Very high risk of cancer in familial Peutz-Jeghers syndrome. Gastroenterology 119:1447–1453 47. Giovannucci E, Harlan DM, Archer MC, Bergenstal RM, Gapstur SM, Habel LA et al (2010) Diabetes and cancer. Diabetes Care 33:1674–1685. https://doi.org/10.2337/dc10-0666 48. Girard C, Duprat F, Terrenoire C, Tinel N, Fosset M, Romey G et al (2001) Genomic and functional characteristics of novel human pancreatic 2P domain K(+) channels. Biochem Biophys Res Commun 282:249–256. https://doi.org/10.1006/bbrc.2001.4562 49. Gloyn AL, Pearson ER, Antcliff JF, Proks P, Bruining GJ, Slingerland AS et al (2004) Activating mutations in the gene encoding the ATP-sensitive Potassium-Channel subunit Kir6.2 and permanent neonatal diabetes. N Engl J Med 350:1838–1849. https://doi.org/10. 1056/NEJMoa032922 50. Goldstein AM, Fraser MC, Struewing JP, Hussussian CJ, Ranade K, Zametkin DP et al (1995) Increased risk of pancreatic cancer in melanoma-prone kindreds with p16INK4 mutations. N Engl J Med 333:970–974. https://doi.org/10.1056/NEJM199510123331504 51. Grant SFA, Thorleifsson G, Reynisdottir I, Benediktsson R, Manolescu A, Sainz J et al (2006) Variant of transcription factor 7-like 2 (TCF7L2) gene confers risk of type 2 diabetes. Nat Genet 38:320–323. https://doi.org/10.1038/ng1732 52. Green RC, Baggenstoss AH, Sprague RG (1958) Diabetes mellitus in association with primary carcinoma of the pancreas. Diabetes 7:308–311 53. Gullo L (1999) Diabetes and the risk of pancreatic cancer. Ann Oncol Off J Eur Soc Med Oncol 10(Suppl 4):79–81 54. Gupta MK, Vadde R (2019a) Identification and characterization of differentially expressed genes in type 2 diabetes using in silico approach. Comput Biol Chem 79:24. https://doi.org/10. 1016/j.compbiolchem.2019.01.010 55. Gupta MK, Vadde R (2019b) Insights into the structure-function relationship of both wild and mutant zinc transporter ZnT8 in human: a computational structural biology approach. J Biomol Struct Dyn:1–22. https://doi.org/10.1080/07391102.2019.1567391 56. Habib SL, Rojna M (2013) Diabetes and risk of Cancer. Int Sch Res Not 2013:1. https://doi. org/10.1155/2013/583786 57. Han J, Kang D, Kim D (2003) Functional properties of four splice variants of a human pancreatic tandem-pore K+ channel, TALK-1. Am J Physiol-Cell Physiol 285:C529–C538. https://doi.org/10.1152/ajpcell.00601.2002 58. Hanis CL, Boerwinkle E, Chakraborty R, Ellsworth DL, Concannon P, Stirling B et al (1996) A genome-wide search for human non-insulin-dependent (type 2) diabetes genes reveals a major susceptibility locus on chromosome 2. Nat Genet 13:161–166. https://doi.org/10.1038/ ng0696-161 59. Hassan MM, Li D, El-Deeb AS, Wolff RA, Bondy ML, Davila M et al (2008) Association between hepatitis B virus and pancreatic Cancer. J Clin Oncol 26:4557–4562. https://doi.org/ 10.1200/JCO.2008.17.3526 60. Hoesel B, Schmid JA (2013) The complexity of NF-κB signaling in inflammation and cancer. Mol Cancer 12:86. https://doi.org/10.1186/1476-4598-12-86 61. Horikawa Y, Oda N, Cox NJ, Li X, Orho-Melander M, Hara M et al (2000) Genetic variation in the gene encoding calpain-10 is associated with type 2 diabetes mellitus. Nat Genet 26:163–175. https://doi.org/10.1038/79876

3

Diabetes and Pancreatic Cancer: A Bidirectional Relationship Perspective

47

62. Hundal RS, Inzucchi SE (2003) Metformin: new understandings, new uses. Drugs 63:1879–1894. https://doi.org/10.2165/00003495-200363180-00001 63. Isaksson B, Strömmer L, Friess H, Büchler MW, Herrington MK, Wang F et al (2003) Impaired insulin action on phosphatidylinositol 3-kinase activity and glucose transport in skeletal muscle of pancreatic cancer patients. Pancreas 26:173–177 64. Ishikawa ONS (1994) Increased secretion of proinsulin in patients with pancreatic cancer. Int J Pancreatol 16:86 65. Isoda K, Young JL, Zirlik A, MacFarlane LA, Tsuboi N, Gerdes N et al (2006) Metformin inhibits proinflammatory responses and nuclear factor-kappaB in human vascular wall cells. Arterioscler Thromb Vasc Biol 26:611–617. https://doi.org/10.1161/01.ATV.0000201938. 78044.75 66. Ji BT, Chow WH, Hsing AW, McLaughlin JK, Dai Q, Gao YT et al (1997) Green tea consumption and the risk of pancreatic and colorectal cancers. Int J Cancer 70:255–258 67. Kamisawa T, Wood LD, Itoi T, Takaori K (2016) Pancreatic cancer. Lancet 388:73–85. https://doi.org/10.1016/S0140-6736(16)00141-0 68. Kang D, Kim D (2004) Single-channel properties and pH sensitivity of two-pore domain K+ channels of the TALK family. Biochem Biophys Res Commun 315:836–844. https://doi.org/ 10.1016/j.bbrc.2004.01.137 69. Karin M (2009) NF-kappaB as a critical link between inflammation and cancer. Cold Spring Harb Perspect Biol 1:a000141. https://doi.org/10.1101/cshperspect.a000141 70. Karlson J, Borg-Karlson AK, Unelius R, Shoshan MC, Wilking N, Ringborg U et al (1996) Inhibition of tumor cell growth by monoterpenes in vitro: evidence of a Ras-independent mechanism of action. Anti-Cancer Drugs 7:422–429 71. Karlson BM, Ekbom A, Josefsson S, McLaughlin JK, Fraumeni JF, Nyrén O (1997) The risk of pancreatic cancer following pancreatitis: an association due to confounding? Gastroenterology 113:587–592 72. Kirpichnikov D, McFarlane SI, Sowers JR (2002) Metformin: an update. Ann Intern Med 137:25–33 73. Koda K, Miyazaki M, Sarashina H, Suwa T, Saito N, Suzuki M et al (2003) A randomized controlled trial of postoperative adjuvant immunochemotherapy for colorectal cancer with oral medicines. Int J Oncol 23:165–172 74. Kornmann M, Maruyama H, Bergmann U, Tangvoranuntakul P, Beger HG, White MF et al (1998) Enhanced expression of the insulin receptor substrate-2 docking protein in human pancreatic cancer. Cancer Res 58:4250–4254 75. Kulie T, Slattengren A, Redmer J, Counts H, Eglash A, Schrager S (2011) Obesity and Women’s health: an evidence-based review. J Am Board Fam Med 24:75–85. https://doi. org/10.3122/jabfm.2011.01.100076 76. Lega IC, Austin PC, Fischer HD, Fung K, Krzyzanowska MK, Amir E et al (2018) The impact of diabetes on breast Cancer treatments and outcomes: a population-based study. Diabetes Care 41:755–761. https://doi.org/10.2337/dc17-2012 77. Leja-Szpak A, Jaworek J, Pierzchalski P, Reiter RJ (2010) Melatonin induces pro-apoptotic signaling pathway in human pancreatic carcinoma cells (PANC-1). J Pineal Res 49:248–255. https://doi.org/10.1111/j.1600-079X.2010.00789.x 78. Lenin M, Ramasamy R, Kulkarani S, Ghose S, Srinivasan ARS, Sathish Babu M (2018) Association of KCNJ11(RS5219) gene polymorphism with biochemical markers of glycemic status and insulin resistance in gestational diabetes mellitus. Meta Gene 16:134–138. https:// doi.org/10.1016/j.mgene.2018.02.003 79. Li D (2012) Diabetes and pancreatic cancer. Mol Carcinog 51:64–74. https://doi.org/10.1002/ mc.20771 80. Liu J, Kazakoff K, Pour PM, Adrian TE (1998) The intracellular mechanism of insulin resistance in the hamster pancreatic ductal adenocarcinoma model. Pancreas 17:359–366

48

M. K. Gupta et al.

81. Liu J, Knezetic JA, Strömmer L, Permert J, Larsson J, Adrian TE (2000) The intracellular mechanism of insulin resistance in pancreatic cancer patients. J Clin Endocrinol Metab 85:1232–1238. https://doi.org/10.1210/jcem.85.3.6400 82. Lowenfels AB, Maisonneuve P (2006) Epidemiology and risk factors for pancreatic cancer. Best Pract Res Clin Gastroenterol 20:197–209. https://doi.org/10.1016/j.bpg.2005.10.001 83. Lowenfels AB, Maisonneuve P, DiMagno EP, Elitsur Y, Gates LK, Perrault J et al (1997) Hereditary pancreatitis and the risk of pancreatic cancer. International hereditary pancreatitis study group. J Natl Cancer Inst 89:442–446 84. Mandalà M, Tondini C (2011) The impact of thromboprophylaxis on cancer survival: focus on pancreatic cancer. Expert Rev Anticancer Ther 11:579–588. https://doi.org/10.1586/era.10. 184 85. Mather KJ, Verma S, Anderson TJ (2001) Improved endothelial function with metformin in type 2 diabetes mellitus. J Am Coll Cardiol 37:1344–1350 86. Monami M, Lamanna C, Balzi D, Marchionni N, Mannucci E (2009) Sulphonylureas and cancer: a case-control study. Acta Diabetol 46:279–284. https://doi.org/10.1007/s00592-0080083-2 87. Morgan AR, Thompson JM, Murphy R, Black PN, Lam W-J, Ferguson LR et al (2010) Obesity and diabetes genes are associated with being born small for gestational age: results from the Auckland birthweight collaborative study. BMC Med Genet 11:125. https://doi.org/ 10.1186/1471-2350-11-125 88. Mössner J, Logsdon CD, Williams JA, Goldfine ID (1985) Insulin, via its own receptor, regulates growth and amylase synthesis in pancreatic acinar AR42J cells. Diabetes 34:891–897 89. Muniraj T, Chari S (2012) Diabetes and pancreatic cancer. Minerva Gastroenterol Dietol 58:331–345 90. Nakchbandi IA, Löhr J-M (2008) Coagulation, anticoagulation and pancreatic carcinoma. Nat Clin Pract Gastroenterol Hepatol 5:445–455. https://doi.org/10.1038/ncpgasthep1184 91. Ndiaye FK, Ortalli A, Canouil M, Huyvaert M, Salazar-Cardozo C, Lecoeur C et al (2017) Expression and functional assessment of candidate type 2 diabetes susceptibility genes identify four new genes contributing to human insulin secretion. Mol Metab 6:459–470. https://doi. org/10.1016/j.molmet.2017.03.011 92. Neuman RJ, Wasson J, Atzmon G, Wainstein J, Yerushalmi Y, Cohen J et al (2010) Genegene interactions lead to higher risk for development of type 2 diabetes in an Ashkenazi Jewish population. PLoS One 5:e9903. https://doi.org/10.1371/journal.pone.0009903 93. Nimptsch K, Rohrmann S, Linseisen J (2008) Dietary intake of vitamin K and risk of prostate cancer in the Heidelberg cohort of the European prospective investigation into Cancer and nutrition (EPIC-Heidelberg). Am J Clin Nutr 87:985–992. https://doi.org/10.1093/ajcn/87.4. 985 94. Novosyadlyy R, LeRoith D (2010) Hyperinsulinemia and type 2 diabetes: impact on cancer. Cell Cycle Georget Tex 9:1449–1450. https://doi.org/10.4161/cc.9.8.11512 95. Odgerel Z, Lee HS, Erdenebileg N, Gandbold S, Luvsanjamba M, Sambuughin N et al (2012) Genetic variants in potassium channels are associated with type 2 diabetes in a Mongolian population. J Diabetes 4:238–242. https://doi.org/10.1111/j.1753-0407.2011.00177.x 96. Ohmura E, Okada M, Onoda N, Kamiya Y, Murakami H, Tsushima T et al (1990) Insulin-like growth factor I and transforming growth factor alpha as autocrine growth factors in human pancreatic cancer cell growth. Cancer Res 50:103–107 97. Ojajärvi IA, Partanen TJ, Ahlbom A, Boffetta P, Hakulinen T, Jourenkova N et al (2000) Occupational exposures and pancreatic cancer: a meta-analysis. Occup Environ Med 57:316–324 98. Pannala R, Leirness JB, Bamlet WR, Basu A, Petersen GM, Chari ST (2008) Prevalence and clinical profile of pancreatic cancer-associated diabetes mellitus. Gastroenterology 134:981–987. https://doi.org/10.1053/j.gastro.2008.01.039

3

Diabetes and Pancreatic Cancer: A Bidirectional Relationship Perspective

49

99. Permert J, Adrian TE, Jacobsson P, Jorfelt L, Fruin AB, Larsson J (1993a) Is profound peripheral insulin resistance in patients with pancreatic cancer caused by a tumor-associated factor? Am J Surg 165:61–66. discussion 66-67 100. Permert J, Ihse I, Jorfeldt L, von Schenck H, Arnqvist HJ, Larsson J (1993b) Pancreatic cancer is associated with impaired glucose metabolism. Eur J Surg Acta Chir 159:101–107 101. Pierce BL, Austin MA, Ahsan H (2011) Association study of type 2 diabetes genetic susceptibility variants and risk of pancreatic cancer: an analysis of PanScan-I data. Cancer Causes Control 22:877–883. https://doi.org/10.1007/s10552-011-9760-5 102. Pisani P (2008) Hyper-insulinaemia and cancer, meta-analyses of epidemiological studies. Arch Physiol Biochem 114:63–70. https://doi.org/10.1080/13813450801954451 103. Porta M, Malats N, Guarner L, Carrato A, Rifà J, Salas A et al (1999) Association between coffee drinking and K-ras mutations in exocrine pancreatic cancer. PANKRAS II study group. J Epidemiol Community Health 53:702–709 104. Prizment A, Gross M, Rasmussen-Torvik L, Peacock J, Anderson K (2012) Genes related to diabetes may be associated with pancreatic cancer in a population-based case-control study in Minnesota. Pancreas 41:50–53. https://doi.org/10.1097/MPA.0b013e3182247625 105. Raderer M, Wrba F, Kornek G, Maca T, Koller DY, Weinlaender G et al (1998) Association between helicobacter pylori infection and pancreatic cancer. Oncology 55:16–19. https://doi. org/10.1159/000011830 106. Rahib L, Smith BD, Aizenberg R, Rosenzweig AB, Fleshman JM, Matrisian LM (2014) 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. https://doi.org/10.1158/ 0008-5472.CAN-14-0155 107. Ramakrishna V, Jailkhani R (2007) Evaluation of oxidative stress in Insulin Dependent Diabetes Mellitus (IDDM) patients. Diagn Pathol 2:22. https://doi.org/10.1186/1746-15962-22 108. Ramakrishna V, Jailkhani R (2008) Oxidative stress in non-insulin-dependent diabetes mellitus (NIDDM) patients. Acta Diabetol 45:41–46. https://doi.org/10.1007/s00592-0070018-3 109. Reid MD, Saka B, Balci S, Goldblum AS, Adsay NV (2014) Molecular genetics of pancreatic neoplasms and their morphologic correlates: an update on recent advances and potential diagnostic applications. Am J Clin Pathol 141:168–180. https://doi.org/10.1309/ AJCP0FKDP7ENVKEV 110. Reinhardt M, Cushman TR, Thearle MS, Krakoff J (2019) Epicardial adipose tissue is a predictor of decreased kidney function and coronary artery calcification in youth- and early adult onset type 2 diabetes mellitus. J Endocrinol Investig 42:979. https://doi.org/10.1007/ s40618-019-1011-8 111. Risch HA (2019) Diabetes and pancreatic cancer: both cause and effect. JNCI J Natl Cancer Inst 111:1–2. https://doi.org/10.1093/jnci/djy093 112. Roberts DL, Dive C, Renehan AG (2010) Biological mechanisms linking obesity and cancer risk: new perspectives. Annu Rev Med 61:301–316. https://doi.org/10.1146/annurev.med. 080708.082713 113. Roberts NJ, Jiao Y, Yu J, Kopelovich L, Petersen GM, Bondy ML et al (2012) ATM mutations in patients with hereditary pancreatic cancer. Cancer Discov 2:41–46. https://doi.org/10.1158/ 2159-8290.CD-11-0194 114. Rothwell PM, Fowkes FGR, Belch JFF, Ogawa H, Warlow CP, Meade TW (2011) Effect of daily aspirin on long-term risk of death due to cancer: analysis of individual patient data from randomised trials. Lancet Lond Engl 377:31–41. https://doi.org/10.1016/S0140-6736(10) 62110-1 115. Sakai K, Imamura M, Tanaka Y, Iwata M, Hirose H, Kaku K et al (2013) Replication study for the association of 9 East Asian GWAS-derived loci with susceptibility to type 2 diabetes in a Japanese population. PLoS One 8:e76317. https://doi.org/10.1371/journal.pone.0076317

50

M. K. Gupta et al.

116. Sale MM, Smith SG, Mychaleckyj JC, Keene KL, Langefeld CD, Leak TS et al (2007) Variants of the transcription factor 7-like 2 (TCF7L2) gene are associated with type 2 diabetes in an African-American population enriched for nephropathy. Diabetes 56:2638–2642. https:// doi.org/10.2337/db07-0012 117. Sanghera DK, Ortega L, Han S, Singh J, Ralhan SK, Wander GS et al (2008) Impact of nine common type 2 diabetes risk polymorphisms in Asian Indian Sikhs: PPARG2 (Pro12Ala), IGF2BP2, TCF7L2 and FTO variants confer a significant risk. BMC Med Genet 9:59. https:// doi.org/10.1186/1471-2350-9-59 118. Schwarts SS, Zeidler A, Moossa AR, Kuku SF, Rubenstein AH (1978) A prospective study of glucose tolerance, insulin, C-peptide, and glucagon responses in patients with pancreatic carcinoma. Am J Dig Dis 23:1107–1114 119. Setiawan VW, Stram DO, Porcel J, Chari ST, Maskarinec G, Le Marchand L et al (2019) Pancreatic Cancer following incident diabetes in African Americans and Latinos: the multiethnic cohort. JNCI J Natl Cancer Inst 111:27–33. https://doi.org/10.1093/jnci/djy090 120. Shah MM, Saif MW (2010) Pancreatic cancer and thrombosis. JOP J Pancreas 11:331–333. https://doi.org/10.6092/1590-8577/3616 121. Siegel RL, Miller KD, Jemal A (2018) Cancer statistics, 2018. CA Cancer J Clin 68:7–30. https://doi.org/10.3322/caac.21442 122. Silverman DT (2001) Risk factors for pancreatic cancer: a case-control study based on direct interviews. Teratog Carcinog Mutagen 21:7–25 123. Society AC (2013) Cancer facts and figures 2013. American Cancer Society Atlanta 124. Someya Y, Tamura Y, Kohmura Y, Aoki K, Kawai S, Daida H et al (2019) A body mass index over 22 kg/m2 at college age is a risk factor for future diabetes in Japanese men. PLoS One 14: e0211067. https://doi.org/10.1371/journal.pone.0211067 125. Stevens RJ, Roddam AW, Beral V (2007) Pancreatic cancer in type 1 and young-onset diabetes: systematic review and meta-analysis. Br J Cancer 96:507–509. https://doi.org/10. 1038/sj.bjc.6603571 126. Stolzenberg-Solomon RZ, Albanes D, Nieto FJ, Hartman TJ, Tangrea JA, Rautalahti M et al (1999) Pancreatic cancer risk and nutrition-related methyl-group availability indicators in male smokers. J Natl Cancer Inst 91:535–541 127. Suzuki H, Li Y, Dong X, Hassan MM, Abbruzzese JL, Li D (2008) Effect of insulin-like growth factor gene polymorphisms alone or in interaction with diabetes on the risk of pancreatic cancer. Cancer Epidemiol Biomark Prev Publ Am Assoc Cancer Res Cosponsored Am Soc Prev Oncol 17:3467–3473. https://doi.org/10.1158/1055-9965.EPI-08-0514 128. Takeda Y, Escribano MJ (1991) Effects of insulin and somatostatin on the growth and the colony formation of two human pancreatic cancer cell lines. J Cancer Res Clin Oncol 117:416–420 129. Thomas PM, Cote GJ, Wohllk N, Haddad B, Mathew PM, Rabl W et al (1995) Mutations in the sulfonylurea receptor gene in familial persistent hyperinsulinemic hypoglycemia of infancy. Science 268:426–429 130. Thomas P, Ye Y, Lightner E (1996) Mutation of the pancreatic islet inward rectifier Kir6.2 also leads to familial persistent hyperinsulinemic hypoglycemia of infancy. Hum Mol Genet 5:1809–1812 131. Vecchia CL, Giordano SH, Hortobagyi GN, Chabner B (2011) Overweight, obesity, diabetes, and risk of breast Cancer: interlocking pieces of the puzzle. Oncologist 16:726–729. https:// doi.org/10.1634/theoncologist.2011-0050 132. Vierra NC, Dadi PK, Jeong I, Dickerson M, Powell DR, Jacobson DA (2015) Type 2 diabetes– associated K+ channel TALK-1 modulates β-cell electrical excitability, second-phase insulin secretion, and glucose homeostasis. Diabetes 64:3818–3828. https://doi.org/10.2337/db150280 133. Vigneri P, Frasca F, Sciacca L, Pandini G, Vigneri R (2009) Diabetes and cancer. Endocr Relat Cancer 16:1103–1123. https://doi.org/10.1677/ERC-09-0087

3

Diabetes and Pancreatic Cancer: A Bidirectional Relationship Perspective

51

134. Vrieling A, Bueno-de-Mesquita HB, Boshuizen HC, Michaud DS, Severinsen MT, Overvad K et al (2010) Cigarette smoking, environmental tobacco smoke exposure and pancreatic cancer risk in the European prospective investigation into Cancer and nutrition. Int J Cancer 126:2394–2403. https://doi.org/10.1002/ijc.24907 135. Wakasugi H, Funakoshi A, Iguchi H (2001) Clinical observations of pancreatic diabetes caused by pancreatic carcinoma, and survival period. Int J Clin Oncol 6:50–54 136. Wang F, Larsson J, Adrian TE, Gasslander T, Permert J (1998) In vitro influences between pancreatic adenocarcinoma cells and pancreatic islets. J Surg Res 79:13–19. https://doi.org/10. 1006/jsre.1998.5393 137. Wang F, Herrington M, Larsson J, Permert J (2003) The relationship between diabetes and pancreatic cancer. Mol Cancer 2:4. https://doi.org/10.1186/1476-4598-2-4 138. Xu C-X, Zhu H-H, Zhu Y-M (2014) Diabetes and cancer: associations, mechanisms, and implications for medical practice. World J Diabetes 5:372–380. https://doi.org/10.4239/wjd. v5.i3.372 139. Yates KR, Welsh J, Echrish HH, Greenman J, Maraveyas A, Madden LA (2011) Pancreatic cancer cell and microparticle procoagulant surface characterization: involvement of membrane-expressed tissue factor, phosphatidylserine and phosphatidylethanolamine. Blood Coagul Fibrinolysis Int J Haemost Thromb 22:680–687. https://doi.org/10.1097/MBC. 0b013e32834ad7bc 140. Zeilhofer HU, Mollenhauer J, Brune K (1989) Selective growth inhibition of ductal pancreatic adenocarcinoma cells by the lysosomotropic agent chloroquine. Cancer Lett 44:61–66 141. Zemskov VS, Procopchuk OL, Susak YM, Zemskov SV, Hodysh YY, Zemskova MV (2000) Ukrain (NSC-631570) in the treatment of pancreas cancer. Drugs Exp Clin Res 26:179–190 142. Zendehdel K, Nyrén O, Östenson C-G, Adami H-O, Ekbom A, Ye W (2003) Cancer incidence in patients with type 1 diabetes mellitus: a population-based cohort study in Sweden. JNCI J Natl Cancer Inst 95:1797–1800. https://doi.org/10.1093/jnci/djg105 143. Zhang P, Li H, Tan X, Chen L, Wang S (2013) Association of metformin use with cancer incidence and mortality: a meta-analysis. Cancer Epidemiol 37:207–218. https://doi.org/10. 1016/j.canep.2012.12.009 144. Zhang Y, Xing Y, Yuan H, Gang X, Guo W, Li Z et al (2018) Impaired glucose metabolisms of patients with obstructive sleep apnea and type 2 diabetes. J Diabetes Res 2018:6714392. https://doi.org/10.1155/2018/6714392 145. Zhu Z, Wang X, Shen Z, Lu Y, Zhong S, Xu C (2013) Risk of bladder cancer in patients with diabetes mellitus: an updated meta-analysis of 36 observational studies. BMC Cancer 13:310. https://doi.org/10.1186/1471-2407-13-310 146. Hu C, Zhang R, Wang C, Wang J, Ma X, Lu J, et al (2009) PPARG, KCNJ11, CDKAL1, CDKN2A-CDKN2B, IDE-KIF11-HHEX, IGF2BP2 and SLC30A8 are associated with type 2 diabetes in a Chinese population. PLoS ONE 4:e7643. https://doi.org/10.1371/journal.pone. 0007643

4

Metabolic Adaptations in Diabetes Mellitus and Cancer Anil Kumar Pasupulati, Nageswara Rao Dunna, and Srikanth Talluri

Abstract

A web of interconnected and flexible metabolic pathways helps in maintaining cellular homeostasis. These flexible metabolic pathways are further deregulated to ensure cell survival during clinical conditions such as diabetes and cancer. In fact, complex metabolic programming is a hallmark of both diabetes and cancer. Normally, glucose absorbed by cells is catabolized through glycolysis to form pyruvate. Pyruvate fuels the citric acid cycle in the mitochondria of aerobic cells. Diabetic condition is presented with a battery of glycolytic abnormalities. Glucose-6-phosphate derived from glucose is converted back to glucose and pumped out of the cell. Similarly, pyruvate is also converted back to glucose. Import of glucose via GLUT4 is also impaired during diabetes. In spite of decelerated glycolysis, gluconeogenesis also occurs in diabetes. In diabetic settings, cells adapt to using acetyl-CoA derived from fatty acids to fuel citric acid cycle. Cancer cells also present with several metabolic abnormalities where glucose-6phosphate powers pentose phosphate pathway and contribute ribose required for the rapid proliferation of transformed cells. Cancer cells convert pyruvate to lactate, which is transported out of cells. Glutamine provides the majority of anaplerotic carbon for the citric acid cycle, and acetyl-CoA derived from citrate contributes to the synthesis of fatty acids. Both 3-phosphoglycerate and pyruvate are spared for the synthesis of amino acids. In the case of diabetes, cells uptake A. K. Pasupulati (*) Department of Biochemistry, School of Life Sciences, University of Hyderabad, Hyderabad, Telangana, India e-mail: [email protected] N. R. Dunna Cancer Genomics Lab, Department of Biotechnology, SASTRA Deemed University, Thanjavur, India S. Talluri Dana-Farber Cancer Institute, Harvard Medical School, Boston, MA, USA # Springer Nature Singapore Pte Ltd. 2019 G. P. Nagaraju, A. BM Reddy (eds.), Exploring Pancreatic Metabolism and Malignancy, https://doi.org/10.1007/978-981-32-9393-9_4

53

54

A. K. Pasupulati et al.

very less glucose compared with normal cells, whereas in case of cancer, cells consume excess glucose due to aerobic glycolysis. In this chapter, we will discuss the major components of the metabolic deregulations in diabetes and cancer. Keywords

Diabetes · Cancer · Metabolism · Randle cycle · Warburg’s effect · Glutamine metabolism

Abbreviations ACC ACLY ACS ACTH AGEs AKT AMPK BRAF CCK CML c-Myc CREBP DAG ETC FAS G6P G6PDH GLP1 GLS1 GLUT4 GOLD GPAT GSH GSSG HIF1α IDH IDO IGFBP3 IKKb IRS LDHA MOLD mTORC NF-kB

Acetyl-CoA carboxylase ATP citrate lyase Acetyl-CoA synthetase Adrenocorticotropic hormone Advanced glycation end products Serine-threonine protein kinase AMP-dependent protein kinase B-Raf proto-oncogene Cholecystokinin Carboxymethyl-lysine Cellular myelocytomatosis cAMP-response element-binding protein Diacylglycerol Electron transport chain Fatty acid synthase Glucose-6-phosphate Glucose-6-phosphate dehydrogenase Glucagon-like peptide-1 Glutaminase-1/L-glutaminase Glucose transporter type 4 Glyoxal-lysine dimer Glycerol-3-phosphate acyltransferase Glutathione Glutathione disulfide Hypoxia-inducible factor 1α Isocitrate dehydrogenase Indoleamine 2,3-dioxygenase Insulin-like growth factor-binding protein Inhibitory kB kinase b Insulin receptor substrate Lactate dehydrogenase A Methylglyoxal-lysine dimer Mammalian target of rapamycin complex 1 Nuclear factor-kB

4

Metabolic Adaptations in Diabetes Mellitus and Cancer

α-KG PC PDH PEPCK PFK PHGDH PTEN SCD SHMT1 SREBP1 T1D and T2D TDO TIGAR TPH1

4.1

55

α-ketoglutarate Pyruvate carboxylase Pyruvate dehydrogenase Phosphoenolpyruvate carboxykinase Phosphofructokinase Phosphoglycerate dehydrogenase Phosphatase and tensin homolog Stearoyl-CoA desaturase Serine hydroxymethyltransferase Sterol response element-binding protein Type 1 diabetes and type 2 diabetes Tryptophan 2,3-dioxygenase TP53-induced glycolysis and apoptosis regulator Tryptophan hydroxylase 1

Introduction

Metabolism, a central theme in biological chemistry, allows cells and organisms survive by providing metabolites and energy they require for maintenance, growth, and propagation. In the first stage of metabolism, all foods are fragmented into their monomers, mainly glucose, amino acids, and fatty acids, which are further subjected to complete breakdown into water and carbon dioxide via pyruvate and acetyl-CoA. Both pyruvate and acetyl-CoA serve as crucial hubs of intermediary metabolism. Transient storage of glucose in its polymeric form as glycogen helps in maintaining optimum glucose levels. Additional glucose can also be synthesized from pyruvate, which is derived from amino acids. Fatty acids that are either absorbed from diet or released from adipose tissue are converted to ketone bodies via acetyl-CoA. The central theme of intermediary metabolism is ensuring tight control of blood glucose levels that are about 100 mg/dl or 5.5 mM. The metabolic pathways diverge into two types: catabolic, those that degrade biomolecules to release free energy, and anabolic, those that polymerize monomers into macromolecules. A great portion of the biopolymers is broken down during catabolism to ensure ATP production. ATP helps meet the energy requirements of cells and also facilitate anabolic pathways. Anabolic pathways ensure the synthesis of macromolecules including fatty acids, proteins, and nucleic acids. Depending on the context, a few metabolic pathways can serve both as catabolic and anabolic. An example for such amphibolic pathway is citric acid cycle, which participates in the synthesis of amino acids in addition to break down of acetyl-CoA. Another example for amphibolic pathway is pentose phosphate pathway.

56

4.2

A. K. Pasupulati et al.

Healthy Individuals Maintain Strict Normoglycemia

Tight regulation of blood glucose levels within a narrow range is referred to as glucose homeostasis. Glucose homeostasis (4–6 mM) is strictly regulated predominantly by the tug-of-war between insulin (that lowers blood glucose) and glucagon (that raises blood glucose) [1]. Classically, insulin and glucagon are considered to regulate blood glucose, but recent studies added several hormonal and neural glucoregulatory factors that influence blood glucose levels moderately (Table 4.1). Insulin and glucagon interact with tissues such as the liver, gut, brain, adipose, and muscle and act antagonistically to each other in regulating blood glucose levels [2–4]. Circulating glucose, through a series of events, stimulates the pancreatic islet (β) cells to secrete insulin. Insulin elicits its action via its receptors expressed in insulin-sensitive tissues such as the liver, muscle, and adipocytes. Insulin is the principal endocrine mediator that decreases the blood glucose levels by several means: (a) promoting glucose uptake by insulin-sensitive tissues, (b) suppressing hepatic and renal glucose production, and (c) reducing circulating free fatty acids (FFA). Metabolic events regulated by insulin are presented in Table 4.2. Low circulating glucose levels stimulate glucagon from pancreatic islet (α) cells, which elicits counterregulatory action against insulin. Glucagon enhances glucose production in the liver by activating glycogenolysis. The effect of glucose production by the glucagon is temporary. Epinephrine, which is secreted from the adrenal medulla, performs multiple actions in target organs: (a) stimulates glycogenolysis and gluconeogenesis in the liver, (b) induces gluconeogenesis in the kidney, and (c) lowers glucose uptake in the skeletal muscle. Pituitary growth hormone (GH) and cortisol

Table 4.1 List of predominant hormones that are associated with the regulation of glucose homeostasis Hormone Insulin

Effect on blood glucose Reduces

Glucagon Somatostatin Epinephrine

Increases Reduces Increases

Glucagon-like peptide-1 (GLP1) Thyroxin

Reduces

Amylin Cortisol ACTH

Reduces Increases Increases

GH

Increases

Increases

Key mode of action Increase glucose uptake by cells; increase glucose conversion to glycogen or fatty acids Increase glycogenolysis; increase gluconeogenesis Suppress glucagon and insulin secretion Increase glycogenolysis. Enhance fatty acids release from adipocytes Enhance insulin secretion; suppress glucagon secretion Increase glycogenolysis; increase intestinal absorption of sugars Suppress glucagon secretion Increases gluconeogenesis; antagonize insulin action Increase cortisol secretion; enhance fatty acids release from adipocytes Antagonize insulin action

4

Metabolic Adaptations in Diabetes Mellitus and Cancer

57

Table 4.2 Insulin-regulated metabolic events pertinent to carbohydrates, lipids, and proteins Metabolic events regulated by insulin Anabolic effects on carbohydrate metabolism Increase the transport of glucose into insulin-sensitive cells (muscle and adipocytes) Reduce the rate of glucose release by the liver via inhibiting glycogenolysis Reduce the blood glucose by (a) stimulating glucose uptake, (b) stimulating glycolysis, and (c) glycogenesis Inhibit gluconeogenesis Anabolic effects on lipid metabolism Stimulates lipogenesis Facilitates the conversion of fatty acids to triglycerides in the liver Reduces the rate of release of free fatty acids from adipose tissue Protein metabolism Increases the transport of free amino acids into hepatocytes and myocytes Stimulates protein synthesis in myocytes

act as glucose counterregulatory hormones, and both act synergistically. Both these hormones stimulate expression of gluconeogenic enzymes and lower glucose transport [4, 5]. Glucagon-like peptide-1 stimulates insulin secretion and suppresses glucagon secretion. Further, neuropeptides (brain-derived neurotrophic factor, gastrin-releasing peptide), enteroendocrine hormones (gastric inhibitory polypeptide, gastrin, and CCK), adipokines, and myokines partially contribute to glucose homeostasis [6].

4.3

Diabetes Mellitus

Diabetes mellitus (D) is a metabolic disorder with impaired secretion of insulin (type 1 diabetes mellitus (T1D)) and peripheral insulin resistance (type 2 diabetes mellitus (T2D)) resulting in elevated blood glucose levels (hyperglycemia) [7–9]. In T1D (previously known as insulin-dependent or juvenile), insulin secretion is absent due to autoimmune destruction of pancreatic β-cells, which occurs subclinically over months to years. In general, T1D develops in childhood or adolescence. However, it can also develop in adults, which often initially appears to be T2D. In T2D (previously known as non-insulin-dependent or adult-onset), insulin levels are relatively inadequate because target tissues developed insulin resistance. Insulin resistance is manifested in impaired suppression of gluconeogenesis in the liver. Peripheral insulin resistance impairs glucose uptake, and together, hepatic and peripheral insulin resistance results in hyperglycemia. In T2D, early in the disease, insulin levels are very high; however, insulin secretion depletes in the later course of the disease. Symptoms of both T1D and T2D include hyperglycemia in addition to polyuria, polyphagia, and polydipsia. Advanced complications include peripheral neuropathy, nephropathy, and predisposition to infection. In this chapter, the metabolic adaptations will be discussed in hyperglycemia and are relevant to both T1D and T2D.

58

4.4

A. K. Pasupulati et al.

Metabolic Adaptations in Diabetes Mellitus

In this section, metabolic adaptation in diabetes will be discussed considering the conditions such as insulin deficiency and hyperglycemia. Diabetes is generally presented with reduced glucose uptake and glucose clearance by the tissues particularly the muscle and adipose. In addition to perturbations in glucose uptake and metabolism in extrahepatic tissues, deregulation in glucose cycling, suppression of hepatic glycolysis, and increase in hepatic gluconeogenesis contribute to hyperglycemia.

4.4.1

Altered Glucose Cycling in Diabetes

Diabetes is presented with decline in glycolytic flux, which is one of the main reasons for hyperglycemia [10]. Decline in glycolytic flux is explained by increase in glucose-6-phosphate conversion to glucose [11, 12]. Normally, glucose is converted to glucose-6-phosphate immediately after taken up by the liver. But, during diabetes, glucose-6-phosphate gets dephosphorylated and returned to the circulation. The possible reason for the observed effect of glucose flux into circulation is both stimulation of glucose-6-phosphatase activity and decreased glucokinase activity [13]. In decreased insulin, glucagon ratio in diabetes results in the repression of glucokinase system [14]. The rate of phosphorylation of glucose is 30% lower in diabetic liver when compared with healthy liver. Another issue with altered glucose cycling in diabetes is the reduced rate of hepatic glycolysis. Diabetic hepatocytes show a 60% reduction in the rate of glycolysis. The liver recycles a large portion of glycolytically derived pyruvate to glucose during diabetes [13]. This pyruvate to glucose cycling inhibits lactate accumulation. Unlike in normal cells where pyruvate is reduced to lactate in the cytoplasm, in the diabetic liver, glycolytically derived pyruvate is channeled to mitochondria. Mitochondrial import of pyruvate is facilitated by the counterion exchange of ketone bodies particularly acetoacetate. Pyruvate that enters into mitochondria is recycled to glucose via gluconeogenesis pathway, and only about 20% pyruvate is oxidized to CO2. Together, both glucose-6-phosphate and pyruvate cycling to glucose contribute to hyperglycemia. More than three-fourth of glucose taken up by the liver get recycled to glucose in diabetic condition. The increase in glucose cycling in diabetic liver is consequence of increased oxidation of free fatty acids.

4.4.2

The Randle Glucose-Fatty Acid Cycle

The Randle cycle (glucose-fatty acid cycle) explains the competition between glucose and fatty acids for their oxidation and uptake [15, 16]. The Randle cycle is posited to depict impaired glucose metabolism in T2D and insulin resistance. Depending on the availability of the fatty acids, metabolic adaptation in cardiac and skeletal muscle shifts back and forth between carbohydrate and fatty acids as

4

Metabolic Adaptations in Diabetes Mellitus and Cancer

59

oxidative energy source [16]. The Randle cycle explains the mechanisms that are involved in the switch from carbohydrate to fat oxidation and glucose-sparing effect in three possible ways as follows: 1. Increased fat oxidation inhibits glucose utilization; fatty acids are taken up from the plasma and are carried to the mitochondria by fatty acid-binding proteins. The fatty acids are transported into the mitochondria by the carnitine palmitoyltransferase system, where they undergo β-oxidation to yield acetylCoA. Citrate formed from acetyl-CoA inhibits pyruvate dehydrogenase (PDH) and phosphofructokinase, respectively. This leads to a buildup of glucose-6-P and inhibition of hexokinase, resulting in reduced glucose uptake and oxidation (Fig. 4.1).

Fig. 4.1 The Randle cycle explains that free fatty acid (FFA) oxidation yields acyl-CoA, which converts into acetyl-CoA in mitochondria. Acetyl-CoA and citrate inhibit pyruvate dehydrogenase (PDH) and phosphofructokinase (PFK) thus blocking glycolysis. FFA-derived acyl-CoA converts into diacylglycerol (DAG), which by a series of events inhibits insulin signaling and prevents GLUT4 translocation and glucose uptake. Acetyl-CoA formed from mitochondrial oxidation of FFA activates pyruvate carboxylase (PC) and phosphoenolpyruvate carboxykinase (PEPCK) and favors gluconeogenesis

60

A. K. Pasupulati et al.

2. Free fatty acids prevent insulin-mediated glucose transporter 4 (GLUT4) translocation. Normally, insulin induces phosphorylation of insulin receptor substrates (IRS), which eventually results in translocation of GLUT4 to membrane for glucose intake. Free fatty acid metabolites such as ceramides, acyl-CoA, and diacylglycerol (DAG) activate several serine/threonine kinases (protein kinase C, nuclear factor-kB, inhibitory kB kinase b (IKKb)), which phosphorylate IRS, and protein kinase B/Akt thus inhibits insulin signaling. Therefore, free fatty acids impair GLUT4-dependent glucose uptake. 3. Free fatty acids induce gluconeogenesis; mitochondrial oxidation of free fatty acids generates acetyl-CoA, which activates pyruvate carboxylase (PC) and phosphoenolpyruvate carboxykinase (PEPCK), thus promoting gluconeogenesis. Another concern is that peroxisomal oxidation of fatty acids generates acetate. Acetate ensures glucose synthesis via peroxisomal glyoxylate pathway.

4.4.3

Aldose Reductase Pathway

During hyperglycemia, hexokinase that has high affinity for glucose becomes saturated, and the excess glucose is sequestered through the aldose reductase (AR) pathway, also known as the polyol pathway [17]. This is a two-step pathway; in the first step, glucose is reduced to sorbitol by AR (EC 1:1:1:21), and this reaction oxidizes NADPH to NADP+ (Fig. 4.2). In the second step, sorbitol is oxidized to fructose, and this reaction is catalyzed by sorbitol dehydrogenase, which also produces NADH from NAD+. The rate of sorbitol conversion to fructose is slower than the rate at which glucose is converted to sorbitol. Therefore, AR pathway results in the accumulation of sorbitol and depletion of NADPH. Depletion of NADPH results in decreased levels of reduced glutathione, a cellular antioxidant. AR

Fig. 4.2 Aldose reductase reduces glucose to sorbitol. Sorbitol dehydrogenase catalyzes sorbitol conversion to fructose. The depletion of NADPH in this pathway results in the NADPH scarcity, which results in decreased synthesis of reduced glutathione, an antioxidant

4

Metabolic Adaptations in Diabetes Mellitus and Cancer

61

pathway is implicated in several complications of diabetes, especially in microvascular damage in the retina, kidney, and nerves [18, 19]. Sorbitol accumulates and produces osmotic stresses on cells by draining water into tissues. Alternatively, fructose formed during polyol pathway contributes to the formation of advanced glycation end products (AGEs). In normoglycemic conditions, polyol pathway is inactive as AR has high Km and low affinity for glucose. The biological significance of polyol pathway is not known; however, fructose has importance in production of spermatocytes.

4.4.4

Nonenzymatic Glycation

The events of nonenzymatic glycation (NEG) involve the irreversible attachment of reducing sugars with free amino groups (predominantly lysine or arginine) of proteins [20]. NEG reactions occur at a slow rate during normal aging; however, these reactions accelerate in diabetic subjects due to the abundance of glucose and fructose. NEG comprises sequel of events that trigger with formation of Schiff’s base between reducing sugar and amino group and intermediate Amadori product that further undergo Maillard reactions to yield advanced glycation end products (AGEs). These AGEs form adduct on afflicted proteins and include large number of heterogeneous chemical structures with cross-linking propensity (Table 4.3). AGEs are fluorescent and relatively insoluble and appear yellow-brown. Carboxymethyl-lysine (CML), pentosidine, argypyrimidine, pyrraline, carboxyethyl-lysine, fructoselysine, and methylglyoxal-derived hydroimidazolones are well-characterized AGEs [20]. Both biochemical and immunohistochemical evidence suggest that CML is predominant AGE in majority of tissues from diabetic patients [20]. NEG of client proteins adversely affects their structural and functional properties [20]. In addition to affecting the physicochemical properties of the afflicted proteins, NEG induces aberrant cellular signaling and alters gene expression profiles [20–23]. HbA1c, a predominant form of glycated hemoglobin, is an index of nonenzymatic glycation in diabetic patients. In addition to the secondary

Table 4.3 Commonly observed AGEs in the biological systems AGEs Carboxymethyl-lysine Pentosidine Argypyrimidine Imidazolone Carboxyethyl lysine GOLD MOLD

Carbohydrate source Glucose, threose Ribose Methylglyoxal 3-Deoxyglucosone Methylglyoxal Glyoxal Methylglyoxal

Table is modified from Ref. [20] GOLD glyoxal-lysine dimer, MOLD methylglyoxal-lysine dimer

Target amino acid Lysine Lysine + arginine Arginine Arginine Lysine Lysine Lysine

62

A. K. Pasupulati et al.

complications of diabetes, AGEs are implicated in many age-related chronic diseases such as cardiovascular diseases, Alzheimer’s disease, and cancer.

4.5

Metabolic Adaptations in Cancer

The hallmark of cancer is an uncontrolled proliferation of aberrant cells. One aberrant cell typically proliferates up to 109 cancer cells, and to survive and proliferate, cancer cells activate certain metabolic pathways. Altered metabolic pathways in cancer cells are needed to generate enough precursors to meet elevated energy needs, cell maintenance, redox balance, and biosynthesis of macromolecules.

4.5.1

Glycolysis and Warburg Effect

Unlike normal cells, where most of the pyruvate from glycolysis enters the citric acid cycle, in malignant cells, pyruvate converts into lactic acid, even in the presence of oxygen [10]. This explicit conversion of most of the pyruvate to lactate by tumor cells was observed by Otto H. Warburg a decade ago. He showed malignant cells actively use glycolysis for ATP generation, even in the presence of oxygen, and the metabolic shift from oxidative phosphorylation to glycolysis in cancer cells might be due to a respiratory injury. Therefore, instead of 36 ATP from the complete oxidation of one glucose molecule, merely 2 ATP are formed in cancer cells by the conversion of pyruvate into lactate by aerobic glycolysis. Although Warburg proposed the mitochondrial defect to be the culprit for aerobic glycolysis in cancer cells, it is also debated that the observed metabolic derangements were as a result of oncogene activation and tumor suppressor inactivation. Nevertheless, majority of tumors display “glucose addiction” and are presented with increased dependence on glycolysis. Glycolysis provides both ATP and metabolic intermediates essential for cancer cell proliferation. The malignant cells demonstrate great demand for glucose and show a high rate of aerobic glycolysis that offers several advantages as follows: (1) Tumor microenvironment is presented with the hypoxia, and under such conditions of impaired oxidative phosphorylation, glycolysis provides sufficient ATP for cells. (2) The intermediates generated during glycolysis provide precursors for the synthesis of fatty acids and nucleic acids, and this seems to be particularly important for actively proliferating tumor cells (Fig. 4.3). Several intermediates of glycolysis including glucose-6-phosphate and fructose-6-phosphate are channeled via pentose phosphate pathway to generate ribose-5-phosphate and NADPH. Ribose-5-phosphate is building block for nucleic acid synthesis. NADPH is critical for maintaining reduced glutathione levels that in turn regulate cellular redox homeostasis. Furthermore, NADPH also supports lipid synthesis. Therefore, cancer cells adapt to glucose as their predominant source for both energy and key components necessary for nucleic acid and fatty acid synthesis. Metastatic cancer cells accumulate lactate. Lactate

4

Metabolic Adaptations in Diabetes Mellitus and Cancer

63

Fig. 4.3 The major metabolic alterations in cancer cells. 1. Aerobic glycolysis and glucose branching. Glycolytic intermediates transform to ribose-5-phosphate, a precursor for synthesis of nucleic acids. 2. Mitochondrial metabolism is propelled by glutamine. Glutamine converted to glutamate becomes the major carbon source for intermediates of citric acid cycle and particularly for anabolic precursors such as oxaloacetate and citrate. Oxaloacetate converted to aspartate participates in purine synthesis. 3. Citrate, which is derived from glucose or glutamine, converts into acetyl-CoA. Acetyl-CoA and its derivative malonyl-CoA undergo condensation to yield fatty acid palmitate

indirectly contributes to the invasion of cancer cells by activating extracellular matrix-remodeling enzymes.

4.5.2

Glutamine Metabolism in Cancer Cells

Glutamine is an indispensable metabolic substrate for cancer cells wherein it acts as a building block via anabolism and serves as an energy source through catabolism (Fig. 4.3). Glutaminase 1 (GLS1) converts glutamine to glutamate. Glutamate conversion to α-ketoglutarate (α-KG) and routing into the citric acid cycle appear to be crucial for the oncogene-induced tumor growth. α-KG is the main source of

64

A. K. Pasupulati et al.

oxaloacetate and malate for fatty acid synthesis in malignant cells. Studies have shown that the expression of GLS1 correlates with growth of the tumor. It is noteworthy that inhibition of glutaminase suppresses growth of tumor. The precise role of glutamine in tumorigenesis remains to be characterized despite the fact that suppression of glutaminase inhibits oncogenic transformation. Glutamine provides the majority of anaplerotic carbon for the citric acid cycle. Further, glutamine contributes to the antioxidant pool in tumor cells. Malate derived from glutamine (via citric acid cycle) serves as a substrate for the malic enzyme. This enzymatic conversion yields NADPH, which is important for the conversion of glutathione (GSH), a cellular antioxidant from its oxidized form glutathione disulfide (GSSG). It was revealed that oncogene c-Myc contributes to glutamine metabolism directly and indirectly. c-Myc stimulates expression of glutamine transporters and helps glutamine import into cells. Alternatively, c-Myc also stimulates GLS1 expression by suppressing the miRNA (miR23A and miR23B), which otherwise represses GLS1 expression [24, 25].

4.5.3

Lipid Metabolism in Cancer Cells

Unlike normal cells that rely on lipid intake to produce molecules of membranes and signaling, cancer cells adapt fatty acid synthesis de novo from citrate, glutamine, and NADPH. Carbon units derived from glucose are utilized in the de novo fatty acid biosynthesis. Breakdown of glucose into citrate triggers this anabolic reaction. Citrate is exported from mitochondria to cytosol and is converted into acetyl-CoA by ATP citrate lyase (ACLY). Irreversible carboxylation of acetyl-CoA produces malonyl-CoA, and this reaction is catalyzed by acetyl-CoA carboxylase (ACC). These two precursors (acetyl-CoA and malonyl-CoA) are repeatedly assembled to yield palmitate (Fig. 4.3). Palmitate is a key source for mono- and polyunsaturated fatty acids, and stearoyl-CoA desaturase (SCD) converts palmitate into palmitoleic acid, which is a monounsaturated fatty acid. During hypoxic conditions, citrate is synthesized from glutamine, thus contributing to lipid synthesis. It is worth mentioning that in rapidly growing tumors, hypoxia prevails, and during hypoxia, SCD activity is significantly reduced. This could explain the reason for decreased amount of glutamine-derived unsaturated fatty acids. To overcome SCD deregulation in hypoxia, cancer cells rely on extracellular unsaturated lysophospholipids [26]. It is interesting to note that aggressive tumors depend on lipids obtained from exogenous sources for their survival, and particularly, metastatic cancer cells skew toward adipocytes [27]. NADPH is another indispensable source required for fatty acid biosynthesis. Cancer cell couples fatty acid oxidation with fatty acid biosynthesis for maintaining NADPH levels. In conditions of nutrient depletion and stress, malignant cells maintain NADPH levels both by reducing fatty acid biosynthesis, where NADPH is consumed, and increasing fatty acid oxidation, where NADPH is generated. It is suggested that these metabolic rearrangements in cancer cells are orchestrated by the AMP-dependent protein kinase (AMPK), which allows survival of cancer cells

4

Metabolic Adaptations in Diabetes Mellitus and Cancer

65

during anchorage-independent growth and nutrient depletion [28]. It is interesting to note that either inhibition of fatty acid oxidation or de novo lipid biosynthesis similarly sensitizes leukemia cells to apoptosis suggesting that these two metabolic events might share common substrates and these two events occur simultaneously [29]. Several enzymes that facilitate lipid metabolism such as ACL, ACC, FAS, and SCD are under transcriptional control by sterol response element-binding protein (SREBP1). This lipogenic transcription factor SREBP1 is a key downstream target of oncogenic BRAF signaling. Sustained lipogenesis through the maintenance of active SREBP1 contributes to therapeutic resistance in BRAF-mutant melanoma [30].

4.5.4

Protein Metabolism in Cancer Cells

The synthesis of proteins is a highly complex and synchronized process and requires involvement of all the amino acids. The rate of protein synthesis is high in cancer cells that increase the overall need for amino acids. Malignant cells have a high requirement for particularly serine and glycine owing to their involvement in nucleotide biosynthesis. mTORC1 is a major signaling pathway involved in protein translation and amino acid metabolism and is activated by multiple oncogenic insults, particularly oncogene c-Myc. Cancer cells are influenced by growth factor signaling to express the amino acid transporters on the surface that facilitate their uptake from surrounding media. Amino acid availability stimulates mTORC1, which in turn stimulates the synthesis of proteins by effecting translation and biogenesis of ribosomes. As discussed above, glutamine uptake is another important trait of cancer cells. Glutamine uptake and the activity of glutaminase enzyme are both triggered by mTORC1, thus providing glutamate for transamination reactions and for the upkeep of the citric acid cycle. Citrate acid cycle intermediates further contribute for the synthesis of various amino acids. When there is a glutamine surplus, it is often exchanged for essential amino acids to trigger mTORC1 and protein biosynthesis. Thus, an abundance of glutamine and essential amino acids empowers mTORC1-mediated activation of protein biosynthesis.

4.6

Factors Responsible for Metabolic Adaptations in Cancer Cells

4.6.1

Hypoxia

It is very well known that cancer cells are often presented with hypoxic conditions, and hypoxia is one of the main culprits contributing to alterations in tumor metabolism. Hypoxia is an important contributor to intra- and inter-tumor cell diversity and stimulates pronounced but nonuniform expression of hypoxia-driven genes in solid tumors [31]. Majority of glycolytic enzymes are upregulated during hypoxia and provide the rationale for increased rate of glycolysis in cancer cells. In contrast,

66

A. K. Pasupulati et al.

expression of lactate dehydrogenase, which converts pyruvate to lactate, increases during hypoxia. Hypoxia-inducible factor 1α (HIF1α) and HIF2α are the two major transcription factors that transduce hypoxic effects on cells. Of the two, HIF1α is considered as the master sensor that coordinates cellular responses in relation with oxygen homeostasis. HIF1α stimulates 9 of the 10 genes involved in glucose uptake and glycolysis including GLUT1, hexokinase, and lactate dehydrogenase [32]. HIF1α induces pyruvate dehydrogenase kinase, which in turn blocks pyruvate dehydrogenase, thereby inhibiting pyruvate conversion to acetyl-CoA. HIF2α elicits cell multiplication during hypoxic conditions via surging c-Myc function. The targets of c-Myc include enzymes of the glycolytic pathway, glutaminolysis, and fatty acid synthesis. It also mediates the overexpression of the glutamine transporters in the membrane and mitochondrial glutaminase enzyme.

4.6.2

Inactivation of Tumor Suppressors and Activation of Oncogenes

The predominant event in cancer is the occurrence of inactivating mutations in tumor suppressor genes and activating mutations in oncogenes. Most cancers are presented with mutations in Tp53 tumor suppressor gene. Anabolic pathways branching off from glycolysis are regulated by p53. PARK2 (Parkinson’s disease-associated gene) is a p53 target and acts as a tumor suppressor gene. PARK2 deficiency ensures cancer cells to adopt the Warburg effect. TP53-induced glycolysis and apoptosis regulator (TIGAR) is an allosteric activator of phosphofructokinase and inhibits glycolysis. Moreover, p53 also suppresses GLUT1 and GLUT4 transcription. Furthermore, p53 reduces cytochrome c oxidase, which helps in assembly of complex IV of electron transport chain (ETC). Therefore, p53 loss in cancer cells aborts ATP production from ETC and shifts the ATP production to glycolysis [33]. Further, p53 regulates mTOR, PTEN, IGFBP3, and AMPK, thus playing a bigger role in affecting tumor metabolism [34]. In most cancer cells, AKT (serine/threonine kinase) is hyperactive. It is considered to be responsible for cancer cells’ addiction to glucose metabolism for survival and proliferation [35]. AKT contributes to glucose metabolism by (1) enhancing GLUT1 transcription, (2) aiding in phosphorylation of glucose by hexokinase, and (3) eliciting expression of glycolytic enzymes through HIF1 and mTOR signaling [32]. Cancer cells are often presented with hyperactive mTOR pathway [36]. mTOR serves as an energy sensor and fosters nutrient uptake, and activation of mTOR increases glycolysis by stimulating HIF1α expression. c-Myc is another oncogene, and its elevated expression is associated with 40% of all cancers [37]. c-Myc affects pathways associated with energy metabolism, particularly glycolysis. c-Myc induces GLUT1, hexokinase, phosphofructokinase, enolase, and lactate dehydrogenase (Table 4.4). As described earlier, c-Myc plays a key role in glutamine metabolism in cancer cells by inducing the expression of glutamine transporters and GLS1 enzyme that converts glutamine to glutamate.

4

Metabolic Adaptations in Diabetes Mellitus and Cancer

67

Table 4.4 List of the oncogenes and transcription factors that regulate metaboic events in cancer cells Oncogenes/ transcription factor HIF1α HIF1α SP1 MYC/HIF1α Lipid metabolism SREBP1

CHO metabolism/aerobic glycolysis PDK1 IDH1/2 G6PDH GLUT1/3, HK2, PFK1/2, LDHA ACLY, ACC, FASN, GPAT, ACS, SCD1

Amino acid metabolism MYC GLUD1/2, GLS2, SHMT1, PHGDH MYC/HIF1α ASS1 NF-kB IDO, TDO CREBP TPH1

4.6.3

Stages of cancer progression Primary tumor cells

Primary tumor cells and proliferating tumor cells Primary tumor cells and proliferating tumor cells

Mitochondrial Dysfunction

Cancer cells are presented with mitochondrial dysfunction resulting in increased levels of succinate, fumarate, and NADPH oxidase. Accumulation of citric acid cycle intermediates shifts energy production to glycolysis. Studies suggest that suppression of the NADPH oxidase gene results in decreased tumor cell proliferation. Mutations in mitochondrial proteins affect events of the citric acid cycle and oxidative phosphorylation. Additionally, p53 upregulates the expression of cytochrome c oxidase 2 and thus surges the mitochondrial respiration rate. In transformed fibroblasts, the morphology and function of the mitochondria are affected, and oxidative phosphorylation is attenuated due to reduced expression of respiratory complex I proteins [38, 39]. Cancer cells possess low mitochondrial respiration and compensate with an increased dependency on glycolysis. It is interesting to note that mutations in mitochondrial DNA occur at higher frequency when compared with normal cells. Predominantly, mutations occur in genes encoding mitochondrial ETC complex I, III, IV, and V, tRNAs, and rRNAs [32]. Possible reason that explains, at least in part, for higher frequency of mutations in mitochondrial DNA is increased production of free radicals in mitochondria of cancer cells [40].

4.7

Summary

It is now accepted that reprograming of metabolic pathways is a hallmark of both diabetes and cancer. In diabetes, it is more clear that either insulin resistance or absolute deficiency of insulin is major culprit in eliciting metabolic reprograming. In

68

A. K. Pasupulati et al.

case of cancer, it is not yet clear what comes first between metabolic reprograming and uncontrolled proliferation. It is intriguing because mutations in genes whose products encode metabolic enzymes are associated with certain cancers. On the other hand, mutations in tumor suppresser genes and oncogenes are directly attributed with some cancers. Studies are being focused on metabolic alterations in cancer cells with an aim to better understand the biology and improve therapeutic responses. Glycolysis (lactate dehydrogenase/hexokinase), mitochondrial metabolism (ETC), and glutamine metabolism (glutaminase) have emerged as therapeutic targets in cancer cells. It is interesting to note that metformin originally considered as antidiabetic drug also exhibits anticancer properties by inhibiting ETC complex and thus mitochondrial ATP production. Cells demonstrate contrasting phenotype pertinent to glucose uptake in case of diabetes and cancer. Cells uptake very less glucose in diabetes, whereas cancer cells consume excess glucose. Nevertheless, it appears that controlling strict hyperglycemia appears to be a common strategy to combat both complications of both diabetes and cancer. Acknowledgments Authors thank Rajkishor Nishad, Lakshmi Prasanna, and Deepti Nabariya for their assistance during the preparation of this chapter.

References 1. Roder PV et al (2016) Pancreatic regulation of glucose homeostasis. Exp Mol Med 48:e219 2. Kahn CR (1985) The molecular mechanism of insulin action. Annu Rev Med 36:429–451 3. Scott RV, Bloom SR (2018) Problem or solution: the strange story of glucagon. Peptides 100:36–41 4. Habegger KM et al (2010) The metabolic actions of glucagon revisited. Nat Rev Endocrinol 6 (12):689–697 5. Sherwin RS, Sacca L (1984) Effect of epinephrine on glucose metabolism in humans: contribution of the liver. Am J Phys 247(2 Pt 1):E157–E165 6. Aronoff SL et al (2004) Glucose metabolism and regulation: beyond insulin and glucagon. Diabetes Spectr 17(3):183–190 7. Blair M (2016) Diabetes mellitus review. Urol Nurs 36(1):27–36 8. Kharroubi AT, Darwish HM (2015) Diabetes mellitus: the epidemic of the century. World J Diabetes 6(6):850–867 9. Kumar PA, Chitra PS, Reddy GB (2013) Metabolic syndrome and associated chronic kidney diseases: nutritional interventions. Rev Endocr Metab Disord 14(3):273–286 10. Tanner LB et al (2018) Four key steps control glycolytic flux in mammalian cells. Cell Syst 7 (1):49–62 e8 11. Efendic S, Wajngot A, Vranic M (1985) Increased activity of the glucose cycle in the liver: early characteristic of type 2 diabetes. Proc Natl Acad Sci USA 82(9):2965–2969 12. Lickley HL et al (1987) Importance of glucagon in the control of futile cycling as studied in alloxan-diabetic dogs. Diabetologia 30(3):175–182 13. Henly DC, Phillips JW, Berry MN (1996) Suppression of glycolysis is associated with an increase in glucose cycling in hepatocytes from diabetic rats. J Biol Chem 271 (19):11268–11271 14. Barzilai N, Rossetti L (1993) Role of glucokinase and glucose-6-phosphatase in the acute and chronic regulation of hepatic glucose fluxes by insulin. J Biol Chem 268(33):25019–25025

4

Metabolic Adaptations in Diabetes Mellitus and Cancer

69

15. Randle PJ (1998) Regulatory interactions between lipids and carbohydrates: the glucose fatty acid cycle after 35 years. Diabetes Metab Rev 14(4):263–283 16. Hue L, Taegtmeyer H (2009) The Randle cycle revisited: a new head for an old hat. Am J Physiol Endocrinol Metab 297(3):E578–E591 17. Anil Kumar P, Bhanuprakash Reddy G (2007) Focus on molecules: aldose reductase. Exp Eye Res 85(6):739–740 18. Srivastava SK et al (2011) Aldose reductase inhibition suppresses oxidative stress-induced inflammatory disorders. Chem Biol Interact 191(1-3):330–338 19. Suryanarayana P et al (2004) Inhibition of aldose reductase by tannoid principles of Emblica officinalis: implications for the prevention of sugar cataract. Mol Vis 10:148–154 20. Kumar Pasupulati A, Chitra PS, Reddy GB (2016) Advanced glycation end products mediated cellular and molecular events in the pathology of diabetic nephropathy. Biomol Concepts 7 (5-6):293–309 21. Kumar PA et al (2016) Carboxymethyl lysine induces EMT in podocytes through transcription factor ZEB2: implications for podocyte depletion and proteinuria in diabetes mellitus. Arch Biochem Biophys 590:10–19 22. Kumar PA, Kumar MS, Reddy GB (2007) Effect of glycation on alpha-crystallin structure and chaperone-like function. Biochem J 408(2):251–258 23. Muthenna P et al (2014) Effect of cinnamon and its procyanidin-B2 enriched fraction on diabetic nephropathy in rats. Chem Biol Interact 222:68–76 24. Wise DR et al (2008) Myc regulates a transcriptional program that stimulates mitochondrial glutaminolysis and leads to glutamine addiction. Proc Natl Acad Sci USA 105 (48):18782–18787 25. Gao P et al (2009) c-Myc suppression of miR-23a/b enhances mitochondrial glutaminase expression and glutamine metabolism. Nature 458(7239):762–765 26. Kamphorst JJ et al (2013) Hypoxic and Ras-transformed cells support growth by scavenging unsaturated fatty acids from lysophospholipids. Proc Natl Acad Sci USA 110(22):8882–8887 27. Nieman KM et al (2011) Adipocytes promote ovarian cancer metastasis and provide energy for rapid tumor growth. Nat Med 17(11):1498–1503 28. Jeon SM, Chandel NS, Hay N (2012) AMPK regulates NADPH homeostasis to promote tumour cell survival during energy stress. Nature 485(7400):661–665 29. Samudio I et al (2010) Pharmacologic inhibition of fatty acid oxidation sensitizes human leukemia cells to apoptosis induction. J Clin Invest 120(1):142–156 30. Talebi A et al (2018) Sustained SREBP-1-dependent lipogenesis as a key mediator of resistance to BRAF-targeted therapy. Nat Commun 9(1):2500 31. Semenza GL (2003) Targeting HIF-1 for cancer therapy. Nat Rev Cancer 3(10):721–732 32. Hammoudi N et al (2011) Metabolic alterations in cancer cells and therapeutic implications. Chin J Cancer 30(8):508–525 33. Matoba S et al (2006) p53 regulates mitochondrial respiration. Science 312(5780):1650–1653 34. Feng Z, Levine AJ (2010) The regulation of energy metabolism and the IGF-1/mTOR pathways by the p53 protein. Trends Cell Biol 20(7):427–434 35. Elstrom RL et al (2004) Akt stimulates aerobic glycolysis in cancer cells. Cancer Res 64 (11):3892–3899 36. Chapuis N et al (2010) Perspectives on inhibiting mTOR as a future treatment strategy for hematological malignancies. Leukemia 24(10):1686–1699 37. Wokolorczyk D et al (2008) A range of cancers is associated with the rs6983267 marker on chromosome 8. Cancer Res 68(23):9982–9986 38. Chiaradonna F et al (2006) Expression of transforming K-Ras oncogene affects mitochondrial function and morphology in mouse fibroblasts. Biochim Biophys Acta 1757(9-10):1338–1356 39. Baracca A et al (2010) Mitochondrial Complex I decrease is responsible for bioenergetic dysfunction in K-ras transformed cells. Biochim Biophys Acta 1797(2):314–323 40. Carew JS, Huang P (2002) Mitochondrial defects in cancer. Mol Cancer 1:9

5

Role of Mitochondria in Pancreatic Metabolism, Diabetes, and Cancer Noble Kumar Talari, Ushodaya Mattam, and Naresh Babu V. Sepuri

Abstract

Ever since from the discovery of mitochondria, it attracted the field of medicine due to its involvement in numerous aspects of cellular metabolism. Mitochondria, the powerhouse of the cell, are extremely important for maintenance of several vital processes such as TCA cycle, generation of cellular energy source adenosine triphosphate (ATP), cell growth, cell death, and signal transduction. Therefore, it is evident that any type of mitochondrial perturbations would result in myriad of diseases. Mitochondrial diseases are also results from nuclear DNA mutations because most proteins involved in mitochondrial metabolism and mitochondrial DNA maintenance are nuclear encoded. Mitochondrial DNA variations are also observed in aging, diabetes, cancer, and neurological diseases such as Parkinson’s and Alzheimer’s diseases. So far, the research accentuated the relationship between mitochondrial dysfunction and multitude of diseases. Pancreas is one such organ that is frequently affected by mitochondrial perturbations. Over the past decade, significant amount of research has been done on mitochondria in pancreatic dysfunction. Heretofore, accumulating evidences suggest that pancreatitis, diabetes, and pancreatic cancer have highlighted the relationship with mitochondrial dysfunction. In this book chapter, we explore the advances that have been made toward identifying the mitochondria as therapeutic target in pancreatic malignancies including pancreatic metabolism, diabetes, and cancer. Keywords

Pancreatic ductal adenocarcinoma · Mitochondria · Type 2 diabetes mellitus · Pancreatic β cells · Insulin and metformin

Co-correspond author: Noble Kumar Talari. N. K. Talari (*) · U. Mattam · N. B. V. Sepuri (*) Department of Biochemistry, School of Life Sciences, University of Hyderabad, Hyderabad, India e-mail: [email protected] # Springer Nature Singapore Pte Ltd. 2019 G. P. Nagaraju, A. BM Reddy (eds.), Exploring Pancreatic Metabolism and Malignancy, https://doi.org/10.1007/978-981-32-9393-9_5

71

72

N. K. Talari et al.

Abbreviations αKG β-Lap Ala AM ARE ASK1 Asp BPTES Cit Cys DON ER Fum GCL GEM Gln Glu GLUD1 Gly GOT GS HA HBP HIF IGF-1 Iso Lac LDHA Mal MDH Met MMP NOX NQO1 NR5A2 NRF2 OAA OAA PDAC PPP PTEN Pyr

Alpha-ketoglutarate β-Lapachone Alanine Adrenomedullin Antioxidant response element Akt-stimulated phosphorylation of apoptosis signal-regulating kinase 1 Aspartate Bis-2-(5-phenylacetamido-1,3,4-thaidazole-2-yl) ethyl sulfide Citrate Cysteine 6-Diazo-5-oxo-L-norleucine Endoplasmic reticulum Fumarate Glutamate cysteine ligase Gemcitabine Glutamine Glutamate Glutamate dehydrogenase Glycine Aspartate transaminase Glutathione synthase Hyaluronan or hyaluronic acid Hexosamine biosynthetic pathway Hypoxia inducible factors Insulin-responsive growth factor Isocitrate Lactate Lactate dehydrogenase A Malate Malate dehydrogenase Metformin Matrix metallo proteinase NADH oxidase NADPH-quinone oxido-reductase 1 Nuclear receptor 5 A2 Nuclear factor erythroid 2-related factor 2 Oxaloacetate Oxaloacetic acid Pancreatic ductal adenocarcinoma Pentose phosphate pathway Phosphatase and tensin homolog Pyruvate

5

Role of Mitochondria in Pancreatic Metabolism, Diabetes, and Cancer

R5P ROS Suc T2DM TCA UCP2 UDP-GlcNAC

5.1

73

Ribose 5-phosphate Reactive oxygen species Succinate Type2 diabetes mellitus Tricarboxylic acid cycle Uncoupling protein 2 Uridine diphosphate N-acetylglucosamine

Introduction

Pancreatic ductal adenocarcinoma (PDAC) is one of the aggressive pancreatic malignancies with increased incidence over the years and becomes the second leading cause of cancer-related death. The 5-year survival rate of PDAC is 8% which pinpoints the high mortality rate of this cancer. The high mortality rate results from several factors such as challenges associated with surgical excision, lack of early diagnostic and biomarkers, and delay in occurrence of clinical symptoms. Hence, 80% of pancreatic cancers are at advanced stage during diagnosis. In addition, limited efficacy of therapeutic treatment options also hampers the outcome of the disease. Therefore, new therapeutic treatment options are required to manage the disease. In recent years, a new phenomenon has been observed that certain cancer cells depend on metabolism for survival and proliferation, which is defined as metabolic addiction. This non-oncogenic addiction is tissue specific and influenced by both genetic and environmental factors. Evidences showed that pancreatic cancer cells depend more on metabolic addiction, and therefore, it is a new area for therapeutic intervention. Pancreatic cancers consist of a dense fibroblastic stroma rich in fibroblasts and immune cells. The fibroblasts deposit huge amount of extracellular matrix proteins during activation, mainly hyaluronan (HA) (non-sulfated glycosaminoglycan), which results in elevated interstitial pressure of PDAC [50, 87]. Eventually, this creates lack of oxygen and lack of nutrient availability condition. In this condition, cancer cells are challenged to maintain redox, metabolic homeostasis, and macromolecular biosynthesis. Under these circumstances, cancer-stromal cell interactions provide metabolically supportive niche for cancer cell survival. This metabolic adaption can occur by numerous mechanisms such as immune suppression and cross talk mechanisms (by swapping cytokines, metabolites, and growth factors) and further by the physical change in the tumor microenvironment that promotes the tumor progression process [40]. Growing evidences suggest that pancreatic cancer is consequence of diabetes, or diabetes is the consequence of the pancreatic cancer. Therefore, the risk for developing other condition is more as they show bidirectional association. The characteristic features of diabetes such as hyperglycemia and insulin resistance are known to contribute for tumor formation. The induction of pancreatic cancer among diabetes mellitus population is indefinable, and it could be due to metabolic, hormonal, and immunological alterations. Several hypothesized mechanism defines the association between these two conditions such as involvement of insulin, insulin-responsive growth factor (IGF-1), leptin, and adiponectin [22, 69, 85, 86, 113].

74

N. K. Talari et al.

Growing evidences also suggest that pancreatic cancer stimulates the occurrence of diabetes. It has been suggested that diabetes is the early symptom of PDAC as it has been observed with 30% of the PDAC patients [14, 71]. Unlike in type 2 diabetes (T2DM), the glucose intolerance is much worse in pancreatic cancer-induced T2DM. Researchers have found that the manifestation of T2DM among pancreatic cancer patients is likely to be a paraneoplastic observation and is potently induced by several factors such as adrenomedullin (AM); pancreatic cancer-derived S-100A8, a N-terminal peptide; and islet amyloid polypeptides (IAPP) [83] that impair glucose tolerance, thus acting as potential mediators of diabetes mellitus in pancreatic cancer. Additionally, chronic pancreatitis and infectious diseases such as Helicobacter pylori, hepatitis B, and HIV virus are also associated with this PDAC [120]. Therefore, identifying a common target which is important for all these pancreatic malignancies would be beneficial. One such target is mitochondria as it is crucial for the maintenance of several vital processes in cells.

5.2

Mitochondria in Pancreatic Malignancies

5.2.1

Mitochondria in Cancer

Since from years, it is known that cancer cells utilize glucose and other building blocks high rate than its counterparts. The discovery of Warburg effect initially leads to the concept that even under oxygen-rich conditions tumor cells tend to shift their metabolism toward glycolysis instead of oxidative phosphorylation. This evidences that mitochondria are either damaged or trivial in cancers. Briefly, it is demonstrated that cancers associate with mutations in mitochondrial DNA, declined mitochondrial copy number, and reduced transcriptional and translation of nuclear genes that encode for mitochondrial proteins [10, 109]. Nonetheless, this perception has been ruled out by the discoveries that oxidative phosphorylation is the ultimate energy source in several malignancies [76]. Furthermore, mitochondria play a pivotal role in cellular metabolism, calcium signaling, and redox signaling. Importantly, in all wellknown hallmarks of cancers such as angiogenesis, resistance to cell death, sustained cellular proliferation, escape from immune response, and deregulated cellular bioenergetics [42, 43], mitochondria play an important role. Hence, mitochondria could be a promising therapeutic target in the treatment of cancers.

5.3

Mitochondrial ROS in Cancer

Mitochondria consist of a double-layered membrane, namely, outer and inner membranes. Mitochondria are the sites of oxidative phosphorylation and respiratory chain complexes present on the inner membrane which facilitates this process. During this process, some of the electrons might escape from electron carriers. So, mitochondria are liable for the production of reactive oxygen species (ROS). This mainly occurred at the sites of mitochondrial electron transport chain (ETC) complexes such as complex I and complex III. The ROS can cause deterioration

5

Role of Mitochondria in Pancreatic Metabolism, Diabetes, and Cancer

75

effect to the mitochondrial DNA. The number of mitochondria varies from cell to cell depending on the bioenergetics of the cell. Each cell has multiple copies of mitochondria, and each mitochondrion contains several copies of mitochondrial DNA. Mitochondria have their own DNA and encode 37 genes among which 13 genes encode polypeptides required for oxidative phosphorylation, 2 genes encode rRNAs, and the remaining 22 genes encode tRNAs. Mitochondrial DNA is more prone for mutations compared to nuclear DNA as it is exposed to ROS and furthermore the absence of histones in mitochondrial genome also enhances the probability of mutation in mitochondrial DNA [57]. Therefore, mitochondrial DNA variations are responsible for dysregulated oxidative phosphorylation and eventually increase the ROS production [35]. In cells, other sources are also responsible for the production of ROS [45]. To combat the harmful effects associated with ROS production, cells have antioxidant defense mechanisms. The function of each antioxidant enzyme is different, for instance, superoxide dismutases act on superoxide, whereas catalases and peroxiredoxin act on hydrogen peroxide. The enzymatic action of superoxide dismutase converts the superoxide produced in the mitochondria to H2O2, while its conversion to O2 and water is mediated by catalases.

5.4

ROS in Progression of Pancreatic Cancer

The most frequent genetic abnormality associated with PDAC is mutation in KRAS gene. However, targeting KRAS is proven yet challenging, and therefore, a combined strategy of targeting tumor-driven signaling pathways could be a promising target to manage PDAC [54]. The tumor dormant cells which survived with KRAS oncogenic ablation are responsible for tumor relapse or recurrence. These cells have the characteristic feature of stem cells and depend more on OXPHOS. The dense stromal network of PDAC also hampers the delivery of nutrients and oxygen, creating a hypoxic stress condition. In such condition, increased glycolysis and hexosamine biosynthesis pathways facilitate the aggressiveness of PDAC [38]. In addition, oncogenic ablation and metabolic changes also create the circumstances that foster the production of ROS in PDAC. Oncogenic KRAS has shown to induce ROS production in numerous mechanisms. Oncogenic KRAS alters the mitochondrial function by suppressing electron transport chain (ETC) complexes such as complex I and complex III [114] and regulates hypoxia-inducible factors (HIFs) via transferrin receptor (Tfr1). Tfr1 regulates Fe homeostasis and is responsible for the production of ROS. As it is evident that mutant KRAS cells depend more on oxidative phosphorylation, Tfr1 levels have been elevated in several malignancies including pancreatic cancer [51]. The altered efficiency of mitochondrial function results in increased ROS production as well as ROS-driven occurrence of 4-hydroxy-2-nonenal (4-HNE) adduct formation with macromolecules. This inhibits mitochondrial proteins and damages mitochondrial DNA to induce pancreatic cancer initiation [64].

76

N. K. Talari et al.

The mitochondrial respiratory chain complexes as well as NADH oxidase (NOX) enzymes promote the production of ROS via KRAS. NADH oxidases are flavoenzymes and catalyze the O2 to generate ROS, including superoxide radicle (O2 ) and hydrogen peroxide (H2O2). NADH oxidases are activated by small GTPase Rac1, and Rac1 expression is seen high in PDAC cell line [28]. Further KRAS-induced ROS formation is regulated by NOX activation via its cytosolic regulatory subunit, i.e., p47phox [81]. Other several groups have also reported the KRAS-NOX-ROS production in human PDAC which again supports the mechanistic regulation of ROS production is mediated by NOX family members [60, 66, 116]. As mitochondria are important for the production of ROS and most of the NOX enzymes act in mitochondria, various studies have demonstrated the altered mitochondrial function in PDAC. Others reported that 70 mitochondrial proteins were differentially altered by KRAS induction, showing the impact and importance of KRAS on mitochondria and cell metabolism. Among these, NADH dehydrogenase 1 alpha sub-complex assembly factor 1 acts as an important molecule for KRAS-induced mitochondrial dysfunction [111]. Another study has shown that KRAS transformation leads to metabolic adaptation to high glycolytic activity, suppresses mitochondrial respiratory chain complex I activity, and creates ROS stress in cancer cells [47]. Similarly, mitochondrial ROS also drives the formation of pancreatic cancer via growth factor signaling in acinar cells [64]. Others have reported that KRAS-induced tumorigenicity also mediated via ROS generated at the complex III maintains the anchorage-independent growth and cell proliferation. Together, all these evidences describe the importance of mitochondrial metabolism in tumor invasiveness [114].

5.5

ROS-Apoptosis-Pancreatic Cancer

In pancreatic cancer, ROS acts like pro-survival mechanism and inhibits apoptosis. Because, NAD(P)H oxidase stimulates the ROS production, its inhibition via siRNA treatment induces the apoptosis via diminishing the phosphorylation of Akt and ASK1 pathway proteins [75, 108]. Additionally, ROS generated by NADPH oxidase also confers antiapoptotic effect by inhibiting the protein tyrosine phosphatases (PTP). Inhibition of PTP in pancreatic cancer results in enhanced phosphorylation of JAK2 and hampers the apoptosis [60]. It is evident that activation of NF-Kb in cancers confers resistance to apoptosis. Therefore, the role of apoptogenic agent Brucein D (BD) has been studied in pancreatic cancer. This study showed BD triggers ROS production via NADPH oxidase, reduces the activation of NF-Kb signaling pathway, and promotes the activation of p38 mitogen-activated protein kinases (MAPK) signaling pathway [59]. The combined treatment of GEM/cannabinoid in pancreatic cancer is more effective to inhibit the pancreatic tumor growth and triggers autophagy process through ROS-dependent mechanism [26]. As most solid tumors exhibit low levels of antioxidant enzymes, the research work from several groups demonstrated that overexpression of adenovirus-MnSOD construct of mitochondrial superoxide dismutase (MnSOD) suppresses the pancreatic tumor growth. Thus, the inhibition of ROS formation

5

Role of Mitochondria in Pancreatic Metabolism, Diabetes, and Cancer

77

Oncogenic KRAS

HIF NADH Oxidase

ROS

ROS DNA damage

JAK2-STAT1/3

p38MAPK NF-kB

Epithelial mesenchymal transition

Apoptosis

Fig. 5.1 Role of KRAS-induced ROS in pancreatic cancer. The mutant form of KRAS induces ROS by modulating mitochondrial metabolism, hypoxia-inducible factors (HIFs), and NADH oxidase. The altered efficiency of mitochondrial function leads to increased ROS production as well ROS mitochondrial proteins and DNA damage to induce pancreatic cancer initiation. Elevated ROS results in activation of p38 MAPK which initiates epithelial mesenchymal transition. In addition, ROS activates antiapoptosis pathways such as NF-κB and JAK/STAT leading to cancer progression

by superoxide dismutases is beneficial for suppression of pancreatic tumor growth [21, 107, 115] (Fig. 5.1).

5.6

KRAS Alters Glucose and Glutamine Metabolism in Pancreatic Cancer

Although the majority of the PDAC harbors oncogenic KRAS, it drives the tumor development by diverse mechanisms; until now, no effective treatment that targets oncogenic KRAS has reached to the clinic. As RAS inhibitors are unsuccessful, it leads to the understanding that RAS is not an endurable target; however, recent

78

N. K. Talari et al.

studies demonstrate that still this approach may abide to treat KRAS-induced malignancies [73, 98, 105]. Early in the 1900s, a lot of efforts have been put forwarded to develop anti-RAS strategies. One such strategy is inhibiting posttranslational modifications of the RAS by using farnesyltransferase inhibitors (FTIs). Farnesyltransferase transfers the farnesyl group to the RAS protein after translation process, and this step is required for the attachment of RAS to the cell membrane and to transmit signals to initiate cell proliferation. However, the results are disappointing because RAS is alternatively activated by geranylgeranyl transferases (GT). Therefore, researchers have tried with Geranylgeranyl transferase inhibitors (GTIs) and combined therapy of FTIs and GTIs. Conversely, this strategy also failed due to high toxicity effect. As all these results are disappointing, this enforces the discovery of small-molecule inhibitor of phosphodiesterase delta (PDEδ). PDEδ is known as prenyl-binding protein, and it interacts with farnesylated KRAS protein and is important for its localization to plasma membrane (PM). Because KRAS signal transduction chiefly depends on its recruitment toward PM, targeting this may suppress the growth and proliferation of cancer cells. Accumulating evidences also showed that PDE inhibitors are effective therapeutic targets [79, 94]; however, further studies are also required to evaluate its efficacy. Another anti-RAS strategy that targets STK33 kinase is also proven unsuccessful by few research groups [3, 67]. Notwithstanding, with these results, the search for RAS as therapeutic target still continued by researchers. Understanding the fact that cancer cells display altered cellular metabolism, the interest have been shifted toward identifying KRAS-driven potential cellular pathway for therapeutic intervention. Therefore, therapeutic targeting of Warburg effect in pancreatic cancer has gained considerable amount of interest in recent years [91]. Encouraging this, a novel CPI-613, a lipoate analog agent known to inhibit cancer-specific mitochondrial metabolism, has shown to be effective as anticancer drug [68, 74, 80]. Therefore, mitochondria gained attraction as a therapeutic target.

5.7

Metabolic Addiction in Pancreatic Cancer

Cancer cells require adequate energy and building blocks for cellular proliferation. Hence, metabolic pathways are rewired in cancer cells so that the nutrients and energy resources are shifted toward anabolic pathways for their survival. Accumulating evidences suggest that reprogramming of cancer cell metabolism is chiefly controlled by various oncogenic signals, particularly KRAS [17, 18, 61, 90]. Integrated genomic, biochemical, and metabolomics approaches have revealed that KRAS controls multiple metabolic pathways for tumor maintenance in pancreatic cancer. It is evident that glucose transporters and rate-limiting enzymes of glucose metabolism are transcriptionally controlled by KRAS. In pancreatic cancer, in consistent with Warburg effect, the metabolic pathways are reprogrammed, consuming glucose and glutamine more and utilizing them toward biosynthetic pathways. In acceptance with this, in pancreatic cancer, it has been shown that the nitrogen donor for nucleotide biosynthetic pathways chiefly comes from glutamine [121]. The altered glucose metabolism contributes high rates of glycolytic flux,

5

Role of Mitochondria in Pancreatic Metabolism, Diabetes, and Cancer

79

activates hexosamine biosynthetic pathway, and promotes ribose biosynthesis. The pentose phosphate pathway (PPP) generates ribose moiety of DNA and is involved in tumor maintenance of pancreatic cancer. Ribose 5-phosphate is generated from non-oxidative arm of PPP utilized in the biosynthesis of DNA/RNA. Supporting this, KRAS activation has shown to upregulate ribose 5-phosphate isomerase A and ribulose-5-phosphate 3-epimerase, the enzymes of non-oxidative arm of PPP [121]. Importantly, KRAS activation regulates the expression of key enzymes of glycolysis such as HK1, HK2, and PFK [33, 90, 121]. In pancreatic cancer, oncogenic KRAS also increases the transcription of LDHA expression so that the pyruvate is more converted toward lactate [19, 121]. Indeed, the posttranslational modification of LDHA also plays an essential role in pancreatic cancer. LDHA is acetylated at K5 position and inhibits its activity there by negatively regulating pancreatic cancer [123]. Another study also suggested that lactate acts as an alternative fuel to support the proliferation of PDAC [38]. Given the importance of lactate as an alternative fuel, researchers have shown that pancreatic cancer growth is reduced by inhibiting the LDHA expression by RNA interference experiments [123]. Furthermore, Raf/MEK signaling pathway is the major effector pathway of KRAS-induced altered glucose metabolism in pancreatic cancer [121]. Thus, pharmacological inhibition of MEK signaling pathway decreases the KRAS-induced enzymes and their downstream metabolites. Additionally, the promoter region of oncogenic KRAS consists of c-Myc-binding elements, thus suppressing c-Myc expression by RNA interference which results in the suppression of same enzymes which were regulated by KRAS [121]. Together, these results recommend that activation of KRAS effectively alters pancreatic glucose metabolism, and therefore, it is an attractive area to treat pancreatic malignancies. The differences in metabolic dependencies distinguish cancer cells different from its normal counterparts. In addition to glucose, several cancer cells have evolved to rely on glutamine because it is a major source of carbon, precursor for glutathione (a major antioxidant), and source of amino group for amino acids such as glycine, serine aspartate, and alanine. The other remarkable feature of altered metabolism in cancer cells is to accentuate the importance of biosynthetic pathways. Cancer cells utilize glutamine for the synthesis of nucleic acids, proteins, and lipids. Therefore, targeting glutamine metabolism is an alluring area of investigation in cancers. Strikingly, in pancreatic cancer, glucose deprivation has shown minimal effect on growth, and cellular redox state of PDAC suggests that pancreatic cancer cells also depend on alternative mechanisms to maintain redox status. Indeed, withdrawal of glutamine from cells results in altered redox status in PDAC [101]. The mitochondria consume glutamine to generate α-ketoglutarate (α-KG) via glutaminase (GLS1) and glutamate dehydrogenase (GLUD1). This α-KG is used as a fuel in the tricarboxylic acid (TCA) cycle. However, in contrast, PDAC cells use transaminases such as aspartate transaminase (GOT2) for the production of α-KG in mitochondria. Simultaneously, this reaction leads to the production of aspartate from oxaloacetic acid (OAA) in cytoplasm. Further, the cytosolic aspartate transaminase (GOT1) acts on aspartate and generates aspartate back into OAA. The OAA produces malate and pyruvate via by malate dehydrogenase (MDH1) and malic enzyme, respectively. Ultimately, this pathway is essential for the production of

80

N. K. Talari et al.

cytosolic NADPH and maintains redox status in PDAC. Hence, targeting of glutamine metabolism (GOT, MDH, and malic enzyme) suppresses the in vitro and in vivo growth of PDAC [101]. In addition to this pathway, KRAS induces Nrf2dependent ROS detoxification to promote tumorigenesis in pancreatic cancer. In general, ROS are mutagenic and known to promote cancer through oxidation of DNA and subsequent mutations in genes that promote cancer cell proliferation. However, in contrast, ROS detoxification program is also known to promote carcinogenesis in pancreatic cancer [25]. During cellular stress, ROS levels were tightly RSP Pentose phosphate pathway

NAD+

Glycolysis

Ala

UDP-GlcNAc

Asp

Hexosamine Biosynthesis

Glu GLS1

GOT2

Asp

OAA αKG

Mal

GPT1

αKG Glu

Cit TCA Cycle

Fum

GOT1

Iso αKG Suc

Pyr

NADH

OAA

NAD+

MDH1

Glu Cys Gly

Oxphos

ROS

ME1

NADP+

NADPH

GSH

GSSG

H2O2

H2O

GSH biosynthesis

GCL, GS

ARE

ATP

Pyr

Mal

NRF2

Cytosol

GFAT1/2

Extracellular matrix

LAC

LDHA

Gln Glu

Glutamine (Gln)

NAD+

NADH Pyr

Extracellular matrix

HK1/2

GLUT1

Glucose

NADH ATP

Fig. 5.2 Rewiring of glucose and glutamine metabolism in pancreatic cancer. Metabolic pathways are reprogrammed and shifted toward anabolic pathways for pancreatic cancer. Oncogenic KRAS upregulates the glycolysis process via regulation of glucose transporters GLUD1. The altered glucose and glutamine metabolism activates hexosamine biosynthetic pathway, and pentose phosphate pathway generates ribose moiety of DNA. NAD+ for glycolysis maintenance will be generated by conversion of pyruvate to lactate through LDHA. Cancer cells utilize aspartate transaminase (GOT2) instead of GLUD1 to generate α-ketoglutarate in the mitochondria and to fuel TCA cycle. This reaction simultaneously creates aspartate from oxaloacetate (OAA) in cytoplasm. The cytosolic aspartate transaminase (GOT1) acts on aspartate and generates aspartate back into OAA. This OAA is then converted into malate and to pyruvate via malate dehydrogenase (MDH1) and by malic enzyme, respectively. Eventually, this pathway is essential for the production of cytosolic NADPH and thereby GSH levels. On the other hand, increased ROS levels also regulate GSH levels through ROS/NRF2/ARE pathway. NRF2 enhances GSH levels by inducing the expression of GS, GCL, and enzymes involved in GSH synthesis. α-KG alpha-ketoglutarate, Ala alanine, Cit citrate, Cys cysteine, Fum fumarate, Gly glycine, GSH reduced glutathione, GSSH oxidized glutathione, NRF2 nuclear factor erythroid 2-related factor 2, ARE antioxidant response element, GS glutathione synthase, GCL glutamate cysteine ligase, Iso isocitrate, Lac lactate, Mal malate, R5P ribose 5-phosphate, Pyr pyruvate, OAA oxaloacetic acid, Suc succinate, UDP-GlcNAC uridine diphosphate N-acetylglucosamine

5

Role of Mitochondria in Pancreatic Metabolism, Diabetes, and Cancer

81

controlled by detoxification program by antioxidant system that is regulated by transcription factor Nrf2. However, mutant KRAS confers a reduced intracellular environment by constitutively activating this antioxidant program. Nrf2 also generates GSH through a series of reactions. The glutamine (Gln)-derived glutamate (Glu) converts to aspartate (Asp) in the mitochondria, shuttles to the cytosol, and generates NADPH in the cytosol. This NADPH alters the redox homeostasis which enables the cell proliferation via reducing glutathione (GSH) levels. The ability to cope up with oxidative stress and nutrient-poor and hypoxic microenvironment in PDAC utilizes the glutamine to maintain the redox homeostasis, thereby supporting proliferation of PDAC (Fig. 5.2).

5.8

Glutamine Metabolism as Therapeutic Target in Pancreatic Cancer

Several metabolic phenotypes have been observed in PDAC among which is Warburg phenotype which heavily depends on glycolysis producing lactate as final product. Reverse Warburg phenotype utilizes the lactate to fuel the TCA cycle intermediates, eventually less dependent on glucose for proliferation. Glutaminolysis phenotype maintains redox balance by providing TCA cycle intermediates via glutamine. In addition, lipid-dependent phenotype promotes tumor growth by synthesizing lipid from TCA cycle intermediates and further utilizes the energy by lipolysis process [62]. Considering the role of glutamine metabolism in pancreatic cancer proliferation, considerable amount of research has been done to demonstrate its efficacy in therapeutics. 13C-labeling experiments show that the majority of the Asp (aspartate, 50–75%) in pancreatic cancers is derived from glutamine. To test this, initially, researchers have inhibited the glutaminase (GLSi) which hampers the production of glutamate (Glu) in the mitochondria. As GLS is activated by Rho GTPases, researches have inhibited the Rho GTPases by small-molecule inhibitor 968 and found that it inhibits the oncogenic transformation in cancers [110]. Other GLS inhibitors CB-839, bis-2-(5-phenylacetamido1,3,4-thaidazole-2-yl)ethyl sulfide (BPTES), had shown antiproliferative activity in many cancers including pancreatic cancer [7, 37, 118]. Based on the efficacy of CB-839 in pancreatic cell culture system, its efficacy has also been tested in vivo in a treatment-resistant autochthonous mouse model of pancreatic cancer and showed a paradoxical effect. Inhibition of glutaminase by CB-839 has shown no antitumor effect in autochthonous mouse model. Treatment of CD-839 on different orthotopic mouse models such as implanting CD-839-sensitive MPDAC cell line into the organ of tumor origin, i.e., pancreata of nude mice, subcutaneous transplantation of MPDAC cell line into mice, and mice bearing flank tumors has failed to show the decrease in in vivo tumor growth [7]. The ineffectiveness of GLSi by CD-839 is further explained as an adaptive reaction to chronic exposure of GLSi, and multiple compensatory pathways have been activated to sustain the proliferation in pancreatic cancer. This shows the evidence that targeting these pathways in combination with

82

N. K. Talari et al.

GLSi is more effective to enhance the therapeutic strategy [7]. The metabolic plasticity observed in pancreatic cancer utilizes both the glutamine-dependent and glutamine-independent pathways to replenish the nutrients. These results relevantly support previous findings that pancreatic cancer cells obtain energy from multiple fuel sources including autophagy, macropinocytosis, and uptake of free amino acids [20, 102, 119]. Pancreatic cells treated with glutamine and without glutamine affect the expression of several differentially regulated genes essential for JAK/STAT and MAPK signaling pathway. The PPI network analysis has shown that MYC, HSPA5, IL18, and CXCR4 have highest connectivity degree and could play major role in pancreatic cell proliferation [52]. Others have reported that pancreatic cells overexpress MUC1 and replenish the TCA cycle intermediates via glutamine but not glucose. Glucose limitation arrests the cells at G1 phase so that the cells can’t enter into S phase as they can’t synthesize DNA due to the disruption in pyrimidine biosynthesis. MUC1 reprograms glutamine metabolism and provides the important metabolite of pyrimidine nucleotide biosynthesis, i.e., aspartate (Asp), via glutamine [36]. Intracellular pH also regulates cancer cell proliferation. A pH value above 7.2 allows cells to enter into S phase of the cell cycle rapidly and eventually into G2 and M phases. It is also evident that high intracellular pH suppresses the mitotic arrest occurred due to DNA damage. Therefore, in cancers, cell cycle checkpoints are bypassed at high intracellular pH which facilitates the cellular proliferation [63, 88, 112], Putney et al. 2003. Despite, intracellular acidic pH promotes apoptosis by allowing a conformational change in Bax protein (proapoptotic cytosolic protein) and permitting its translocation to mitochondria [55]. Increase in intracellular pH results in low pH in extracellular milieu which further promotes malignancies [32, 82]. At low pH condition, pancreatic cancers rely more on mitochondria in terms of increased glutamine uptake and oxidative phosphorylation. The chronic acidosis stress condition observed due to low pH can be overcome by glutamine metabolism as it serves bioenergetic needs and maintains ROS balance. Aspartate transaminase (GOT1) which converts Asp to OAA is further converted to pyruvate via malic enzyme (ME1). ME1 catalysis produces the NADPH utilized to quench ROS. Thus, anaplerotic glutamine metabolism maintains ROS balance during chronic acidosis condition for the survival of pancreatic cancer [1]. Markedly, even after the improvements in surgical resection, most pancreatic cancer patients will experience recurrence even after complete resection. Therefore, adjuvant systemic therapy helps to overcome the recurrence rate and improves the outcome. Earlier, gemcitabine (GEM) treatment therapy has been preferred as a therapeutic target; however, it fails due to development of innate or adapted drug resistance, eventually leading to poor patient outcomes. Hence, identifying drug or drug combinations which improves the current treatment options greatly benefits the patients. In view of pancreatic cancers’ dependency on glutamine metabolism, in multiple aspects, targeting this pathway may represent an effective therapeutic approach for pancreatic cancers. Accordingly, researchers have used glutamine analogs such as 6-diazo-5-oxo-L-norleucine (DON, interfere with nucleotide and protein biosynthetic pathways). Interestingly, it has been shown that DON reduces

5

Role of Mitochondria in Pancreatic Metabolism, Diabetes, and Cancer

83

the tumor growth and inhibits the metastasis in a mouse model of systemic metastasis [97]. Others have reported that disrupting glutamine metabolism in pancreatic cancers sensitizes the chemoresistant PDAC to GEM therapy. DON treatment to GEM-R pancreatic cell lines effectively reduces the EGFR and AKT signaling pathways, reduces the proliferation, and sensitizes the chemoresistance in PDAC. Indeed, disruption of hexosamine biosynthetic pathway (HBP) via glutamine analogs (DON) also affects glycan biosynthesis and protein glycosylation [15]. In search for glutamine metabolism as an effective therapeutic treatment regime in pancreatic cancers, GLS inhibitors in combination with β-lapachone (β-Lap) were also tested and showed high tumor selectivity. β-Lap is a known targeted cancer therapeutic and forms ROS in tumor cells in an NADPH-quinone oxido-reductase 1 (NQO1)-specific manner. Further, in PDAC, NQO1 is highly elevated, and therefore, it is an attractive target. NQO1 consumes β-Lap as a substrate and produces hydroquinone form of β-Lap. It reacts with oxygen and is converted into its original form by consuming NADPH, eventually generating superoxide molecules. Furthermore, NADPH pools are reduced by treating pancreatic cells with bis-2-(5-phenylacetamido-1,3,4-thaidazole-2-yl) ethyl sulfide BPTES or CB-839 and inhibit the glutamine metabolism (GLSi) along with the tumor-selected ROS formation. Therefore, GLS inhibitors (GLSi) in combination with β-lap are expected to enhance the efficacy for the treatment of PDAC [13].

5.9

Pancreatic Cancer and Diabetes

The association between pancreatic cancer and diabetes has been first published in the year 1995. This study showed that the risk for developing pancreatic cancer is twofold high in diabetes mellitus patients [30]. The second meta-analysis (includes 17 case controls, 19 cohort and nested cohort studies), which has been published in the year 2005, demonstrates that individuals with diabetes have 50% higher chances for developing pancreatic cancer [49]. Few other cohort studies have also determined that manifestation of pancreatic cancer is high in people with diabetes [5, 34, 100]. Another large cohort study (29, 133) revealed that pancreatic cancer malignancy could be predicted by serum insulin levels [103]. All these findings confirm that people who have long-term diabetes are having higher chances for the manifestation of pancreatic cancer. Growing evidences suggest that insulin-like growth factor (IGF) have been associated with increased risk for several cancers. IGF-1 is more potent mitogen than insulin, known to inhibit tumor suppressor phosphatase and tensin homolog (PTEN), and eventually promotes pancreatic cancer proliferation by activating IGF-1/PI3K/Akt signaling pathway [69, 85]. Altered hormonal regulation among adiponectin and leptin (adipose tissue secreted hormones) is known to be associated with pancreatic cancers. Adiponectin hormone circulating levels are inversely proportional to plasma insulin. Thus, low levels of adiponectin increase the risk for developing type 2 diabetes mellitus obesity-related malignancies [22, 113]. Indeed, in vivo studies in pancreatic cancer showed that tumor growth is inversely correlated

84

N. K. Talari et al.

with serum adiponectin levels [124]. A genome-wide analysis also demonstrates that pancreatic-susceptible loci on chromosomes 10p 14 is nuclear receptor 5 A2 (NR5A2), which is essential for transactivation of adiponectin gene. Together, all these findings suggest the possible role of adiponectin in the devolvement of pancreatic cancer. Another hormone leptin which usually associates with obesity and type 2 DM also increases the risk for developing pancreatic cancer [104]. Leptin promotes the pancreatic cancer via JAK/STAT3 signaling pathway by upregulating the expression of matrix metalloproteinase (MMP-13) [31]. Another peptide hormone adrenomedullin (AM) also shown to impair the glucose tolerance by obstructing the insulin secretion in β cells. Hence, AM also acts as a mediator of inducing diabetes [2]. On the other hand, pancreatic cancer-derived S-100A8, an N-terminal peptide [4], and islet amyloid polypeptides (IAPP) [83] and adipose tissue inflammation also impair glucose tolerance, thus acting as potential mediators of diabetes mellitus in pancreatic cancer.

5.10

Metformin as Therapeutic Target in Pancreatic Cancer and Diabetes

Due to the unique relationship of diabetes and pancreatic cancer, antidiabetic drugs that lower insulin resistance and hyperinsulinemia are considered as cancer prevention strategies. One such drug is metformin (Met), a synthetic analog of naturally occurring biguanides. Met is used as standard therapy for diabetes since from 1960s [29] and is recognized as a safe drug. At molecular level, metformin exhibits antiproliferative characteristics in several ways such as it elevates AMP/ATP ratio and activates AMPK/mTOR pathway [8, 84], decreases circulating insulin and blood glucose levels [27], and represses gluconeogenesis by inhibiting the glycerophosphate dehydrogenase of the mitochondria [70], reducing the mitochondrial ETC complex I, thereby tumor respiration [11]. Metformin exits with hydrophobic cation that is responsible for Met uptake by tumor cells. However, metformin lipophilic analogs such as phenformin are taken up by cells via alternative mechanisms [96]. Except from its role in increasing acidosis during diabetic therapy, it has more potential than metformin in inhibiting pancreatic cancer proliferation [58]. Met decreases the ATP production by inhibiting the mitochondrial ETC complex I. This leads to the AMPK activation which in turn disrupts the insulin IGF-1 signaling via mTOR inhibition. Increase in mTOR signaling has been associated with many human pathological conditions such as obesity, type 2 diabetes, and cancer [23, 72, 117]. Therefore, targeting mTOR could be an attractive therapeutic target. Inhibition of mTOR by metformin results in decrease in protein synthesis and cell growth. Interestingly, in the absence of AMPK, metformin and phenformin also inhibit mTOR via Rag GTPases and REDD (mTOR negative regulator) which deciphers the AMPK-independent inhibition of mTOR by metformin and phenformin [53]. Rag GTPases are known to interact with mTORC1 and promotes its recruitment toward its activator Rheb (activator of protein kinase activity of

5

Role of Mitochondria in Pancreatic Metabolism, Diabetes, and Cancer

85

mTORC1) [93]. Metformin increases the REDD expression in p53-dependent manner by inhibiting mTOR and arrests the cell growth [6]. In mice xenografts, Met has shown to reduce the pancreatic cancer growth, when administered orally and intraperitoneally. It also reduces the phosphorylation S6 ribosomal protein and ERK and affects cell growth [56]. Interestingly, at physiological glucose concentration (5 mM), metformin-induced mTOR inhibition in pancreatic cancer is elevated, while it is low at supraphysiological glucose concentrations (25 mM). This highlights the concept that at low glucose levels pancreatic cancer cells rely more on mitochondrial metabolism, and in such condition, they are more sensitive to mitochondrial respiratory inhibitors [92, 99]. Additionally, metformin in combination with 2-deoxyglucose (2-DG, glycolysis inhibitor) enhances the ATP depletion along with antiproliferative effects that have highlighted targeting mitochondrial bioenergetics and glycolysis pathway are productive in the treatment of pancreatic cancer [16].

5.11

Targeting Mitophagy in Pancreatic Cancer and Diabetes: Does the Shoe Fits?

Mitophagy is the selective mode of autophagy that targets the damaged mitochondria to autophagosome and their subsequent catabolism in lysosomes, maintains the mitochondrial integrity and function, and ensures the cellular homeostasis. Mitophagy is induced by several factors such as mitochondrial membrane depolarization, hypoxia, nutrient deprivation, inflammation, and DNA damage. Since mitochondrial homeostasis is an important phenomenon which maintains mitochondrial quality control, the defects in mitophagy have been seen in several pathological situations including cancer and diabetes. Many mitophagy adapter proteins have shown to be altered in such conditions including p62/SQSTM1, BNIP3, BNIPL, PARK2, and FANCC. During mitophagy, the formation of pre-initiation and initiation complex are comparable with general autophagy. Therefore, the defects in mitophagy are also attributed to general autophagy machinery. Strikingly, the nascent phagophore membrane is formed by the contribution from the endoplasmic reticulum (ER) and mitochondria. Hence, mitochondrial integrity is essential to modulate the production of phagophore membranes [9, 41]. Mitochondrial perturbations are vital in the progression of type 2 diabetes. Defects in mitochondria impair insulin signaling through decrease in respiration capacity, ATP production, and high ROS levels. High cytoplasmic ATP levels from glucose oxidation are sensed by pancreatic β cells to secrete insulin. Briefly, in such condition, ATP-sensitive KATP channel is closed causing depolarization of the plasma membrane, calcium influx, and finally insulin secretion. During respiration, uncoupling protein (UCP2) is known to mediate proton leak and decreases the ATP production. Therefore, UCP2 negatively regulates insulin secretion and acts as a mediator of pancreatic β cell dysfunction, obesity, and type 2 diabetes [122]. In addition, damaged mitochondria also increase the apoptosis of pancreatic β cell reducing the mass of β cell in type 2 diabetes [106]. Moreover, the ultrastructural

86

N. K. Talari et al.

analysis of diabetic β cell reveals that mitochondria are compromised in its shape such as round and distorted cristae, and incorporation of mitochondria into autophagic vacuoles suggests the importance of mitochondria in pancreatic β cell function [46, 65]. PINK1 is a serine/threonine protein kinase, and it detects and accumulates in damaged mitochondria. The accumulation of PINK1 recruits the translocation of Parkin (E3 ubiquitin ligase) from cytosol to mitochondria. Further, mitochondrial outer membrane proteins are ubiquitinated by Parkin, and this facilitates the recruitment of mitophagy cargo proteins such as SQSTM1, calcium-binding coiled-coil domain 2 CALCOCO2/NDP52, and optineurin to trigger the mitophagy [77]. PARK6 gene which encodes PINK1 has shown to reduce in obesity and type 2 diabetes implicating its importance in glucose metabolism [89, 95]. Loss of PINK also results in mitochondrial dysfunction in pancreatic β cells and impairs glucose uptake [24]. Additionally, protein folding, aggregation, and accumulation are also crucial for the manifestation of type 2 diabetes. This confers T2DM as protein misfolding disorder (PMD). Pancreatic β cells also secrete amylin protein along with insulin, and its aggregation mediates pancreatic β cell dysfunction and type 2 diabetes [39, 48]. Amylin overexpression in pancreatic β cell has shown to activate mTOR and hampers mitophagy [44]. The levels of chaperone proteins (Bip, protein disulfide isomerase, chemical chaperone such as tauroursodeoxycholic acid (TUDCA) and 4-phenylbutyric acid (4-PBA)) also alter the pancreatic β cell viability and function and improve the glucose homeostasis in T2DM [12, 78]. To sum up, as the defects in mitophagy have been associated with all pancreatic malignancies, targeting mitophagy could be a promising therapeutic target.

5.12

Concluding Remarks

Mitochondria gained a substantial amount of interest in pancreatic malignancies due to its involvement in several vital cellular processes. Predominantly, mitochondrial ROS, mutations in mitochondrial DNA, altered mitochondrial proteins, mitochondrial antioxidant system, respiration, oxidative phosphorylation, defects in mitophagy, metabolic addiction such as rewiring of glucose, and glutamine metabolism have been reported in pancreatic malignancies. Because pancreatic malignancies are bidirectionally association with diabetes, antidiabetic drugs are popular and known to be effective for the treatment options. Therefore, mitochondria as a personalized treatment for pancreatic malignancies are an interesting approach and need to be validated more in the near future. Acknowledgements Authors would like to acknowledge funding from DST-SERB (File number CRG/2018/001028) and DBT (File number BT/PR10319/BRB/10/1267/2013) to Prof. Naresh Babu V Sepuri. Noble Kumar Talari would like to acknowledge CSIR (09/414(1192)/2019EMR-1) for the Research Associate fellowship. Ushodaya Mattam would like to acknowledge DBT (BT/PR19439/BIC/101/431/2016) for the fellowship. The authors also thank their lab members for suggestions.

5

Role of Mitochondria in Pancreatic Metabolism, Diabetes, and Cancer

87

References 1. Abrego J, Gunda V, Vernucci E, Shukla SK, King RJ, Dasgupta A, Goode G, Murthy D, Yu F, Singh PK (2017) GOT1-mediated anaplerotic glutamine metabolism regulates chronic acidosis stress in pancreatic cancer cells. Cancer Lett 400:37–46. https://doi.org/10.1016/j.canlet. 2017.04.029. Epub 2017 Apr 26 2. Aggarwal G, Ramachandran V, Javeed N, Arumugam T, Dutta S, Klee GG, Klee EW, Smyrk TC, Bamlet W, Han JJ, Rumie Vittar NB, de Andrade M, Mukhopadhyay D, Petersen GM, Fernandez-Zapico ME, Logsdon CD, Chari ST (2012) Adrenomedullin is up-regulated in patients with pancreatic cancer and causes insulin resistance in β cells and mice. Gastroenterology 143(6):1510–1517.e1. https://doi.org/10.1053/j.gastro.2012.08.044. Epub 2012 Sep 6 3. Babij C, Zhang Y, Kurzeja RJ, Munzli A, Shehabeldin A, Fernando M, Quon K, Kassner PD, Ruefli-Brasse AA, Watson VJ, Fajardo F, Jackson A, Zondlo J, Sun Y, Ellison AR, Plewa CA, San MT, Robinson J, McCarter J, Schwandner R, Judd T, Carnahan J, Dussault I (2011) STK33 kinase activity is nonessential in KRAS-dependent cancer cells. Cancer Res 71 (17):5818–5826. https://doi.org/10.1158/0008-5472.CAN-11-0778. Epub 2011 Jul 8 4. Basso D, Greco E, Fogar P, Pucci P, Flagiello A, Baldo G, Giunco S, Valerio A, Navaglia F, Zambon CF, Falda A, Pedrazzoli S, Plebani M (2006) Pancreatic cancer-derived S-100A8 Nterminal peptide: a diabetes cause? Clin Chim Acta 372(1–2):120–128. Epub 2006 Mar 29 5. Batty GD, Shipley MJ, Marmot M, Smith GD (2004) Diabetes status and post-load plasma glucose concentration in relation to site-specific cancer mortality: findings from the original Whitehall study. Cancer Causes Control 15(9):873–881 6. Ben Sahra I, Regazzetti C, Robert G, Laurent K, Le Marchand-Brustel Y, Auberger P, Tanti JF, Giorgetti-Peraldi S, Bost F (2011) Metformin, independent of AMPK, induces mTOR inhibition and cell-cycle arrest through REDD1. Cancer Res 71(13):4366–4372. https://doi. org/10.1158/0008-5472.CAN-10-1769. Epub 2011 May 3 7. Biancur DE, Paulo JA, Małachowska B, Quiles Del Rey M, Sousa CM, Wang X, Sohn ASW, Chu GC, Gygi SP, Harper JW, Fendler W, Mancias JD, Kimmelman AC (2017) Compensatory metabolic networks in pancreatic cancers upon perturbation of glutamine metabolism. Nat Commun 8:15965. https://doi.org/10.1038/ncomms15965 8. Birsoy K, Sabatini DM, Possemato R (2012) Untuning the tumor metabolic machine: targeting cancer metabolism: a bedside lesson. Nat Med 18(7):1022–1023. https://doi.org/10.1038/nm. 2870 9. Bockler S (2014) Westermann B Mitochondrial ER contacts are crucial for mitophagy in yeast. Dev Cell 28(4):450–458. https://doi.org/10.1016/j.devcel.2014.01.012. Epub 2014 Feb 13 10. Brandon M, Baldi P, Wallace DC (2006) Mitochondrial mutations in cancer. Oncogene 25 (34):4647–4662 11. Bridges HR, Jones AJ, Pollak MN, Hirst J (2014) Effects of metformin and other biguanides on oxidative phosphorylation in mitochondria. Biochem J 462(3):475–487. https://doi.org/10. 1042/BJ20140620 12. Cadavez L, Montane J, Alcarraz-Vizán G, Visa M, Vidal-Fàbrega L, Servitja JM, Novials A (2014) Chaperones ameliorate beta cell dysfunction associated with human islet amyloid polypeptide overexpression. PLoS One 9(7):e101797. https://doi.org/10.1371/journal.pone. 0101797. eCollection 2014 13. Chakrabarti G, Moore ZR, Luo X, Ilcheva M, Ali A, Padanad M, Zhou Y, Xie Y, Burma S, Scaglioni PP, Cantley LC, DeBerardinis RJ, Kimmelman AC, Lyssiotis CA, Boothman DA (2015) Targeting glutamine metabolism sensitizes pancreatic cancer to PARP-driven metabolic catastrophe induced by ß-lapachone. Cancer Metab 3:12. https://doi.org/10.1186/ s40170-015-0137-1. eCollection 2015 14. Chari ST, Leibson CL, Rabe KG, Timmons LJ, Ransom J, de Andrade M, Petersen GM (2008) Pancreatic cancer-associated diabetes mellitus: prevalence and temporal association with diagnosis of cancer. Gastroenterology 134(1):95–101. Epub 2007 Oct 26

88

N. K. Talari et al.

15. Chen R, Lai LA, Sullivan Y, Wong M, Wang L, Riddell J, Jung L, Pillarisetty VG, Brentnall TA, Pan S (2017) Disrupting glutamine metabolic pathways to sensitize gemcitabine-resistant pancreatic cancer. Sci Rep 7(1):7950. https://doi.org/10.1038/s41598-017-08436-6 16. Cheng G, Zielonka J, McAllister D, Tsai S, Dwinell MB, Kalyanaraman B (2014) Profiling and targeting of cellular bioenergetics: inhibition of pancreatic cancer cell proliferation. Br J Cancer 111(1):85–93. https://doi.org/10.1038/bjc.2014.272. Epub 2014 May 27 17. Chiaradonna F, Magnani C, Sacco E, Manzoni R, Alberghina L, Vanoni M (2005) Acquired glucose sensitivity of k-ras transformed fibroblasts. Biochem Soc Trans 33(Pt 1):297–299 18. Chiaradonna F, Sacco E, Manzoni R, Giorgio M, Vanoni M, Alberghina L (2006) Rasdependent carbon metabolism and transformation in mouse fibroblasts. Oncogene 25 (39):5391–5404. Epub 2006 Apr 10 19. Chun SY, Johnson C, Washburn JG, Cruz-Correa MR, Dang DT, Dang LH (2010) Oncogenic KRAS modulates mitochondrial metabolism in human colon cancer cells by inducing HIF-1α and HIF-2α target genes. Mol Cancer 9:293. https://doi.org/10.1186/1476-4598-9-293 20. Commisso C, Davidson SM, Soydaner-Azeloglu RG, Parker SJ, Kamphorst JJ, Hackett S, Grabocka E, Nofal M, Drebin JA, Thompson CB, Rabinowitz JD, Metallo CM, Vander Heiden MG, Bar-Sagi D (2013) Macropinocytosis of protein is an amino acid supply route in Ras-transformed cells. Nature 497(7451):633–637. https://doi.org/10.1038/nature12138. Epub 2013 May 12 21. Cullen JJ, Weydert C, Hinkhouse MM, Ritchie J, Domann FE, Spitz D, Oberley LW (2003) The role of manganese superoxide dismutase in the growth of pancreatic adenocarcinoma. Cancer Res 63(6):1297–1303 22. Cust AE, Kaaks R, Friedenreich C, Bonnet F, Laville M, Lukanova A, Rinaldi S, Dossus L, Slimani N, Lundin E, Tjønneland A, Olsen A, Overvad K, Clavel-Chapelon F, Mesrine S, Joulin V, Linseisen J, Rohrmann S, Pischon T, Boeing H, Trichopoulos D, Trichopoulou A, Benetou V, Palli D, Berrino F, Tumino R, Sacerdote C, Mattiello A, Quirós JR, Mendez MA, Sánchez MJ, Larrañaga N, Tormo MJ, Ardanaz E, Bueno-de-Mesquita HB, Peeters PH, van Gils CH, Khaw KT, Bingham S, Allen N, Key T, Jenab M, Riboli E (2007) Plasma adiponectin levels and endometrial cancer risk in pre- and postmenopausal women. J Clin Endocrinol Metab 92(1):255–263. Epub 2006 Oct 24 23. Dann SG, Selvaraj A, Thomas G (2007) mTOR Complex1-S6K1 signaling: at the crossroads of obesity, diabetes and cancer. Trends Mol Med 13(6):252–259. Epub 2007 Apr 23 24. Deas E, Piipari K, Machhada A, Li A, Gutierrez-del-Arroyo A, Withers DJ, Wood NW, Abramov AY (2014) PINK1 deficiency in β-cells increases basal insulin secretion and improves glucose tolerance in mice. Open Biol 4:140051. https://doi.org/10.1098/rsob.140051 25. DeNicola GM, Karreth FA, Humpton TJ, Gopinathan A, Wei C, Frese K, Mangal D, Yu KH, Yeo CJ, Calhoun ES, Scrimieri F, Winter JM, Hruban RH, Iacobuzio-Donahue C, Kern SE, Blair IA, Tuveson DA (2011) Oncogene-induced Nrf2 transcription promotes ROS detoxification and tumorigenesis. Nature 475(7354):106–109. https://doi.org/10.1038/nature10189 26. Donadelli M, Dando I, Zaniboni T, Costanzo C, Dalla Pozza E, Scupoli MT, Scarpa A, Zappavigna S, Marra M, Abbruzzese A, Bifulco M, Caraglia M, Palmieri M (2011) Gemcitabine/cannabinoid combination triggers autophagy in pancreatic cancer cells through a ROS-mediated mechanism. Cell Death Dis 2:e152. https://doi.org/10.1038/cddis.2011.36 27. Dowling RJ, Niraula S, Chang MC, Done SJ, Ennis M, McCready DR, Leong WL, Escallon JM, Reedijk M, Goodwin PJ, Stambolic V (2015) Changes in insulin receptor signaling underlie neoadjuvant metformin administration in breast cancer: a prospective window of opportunity neoadjuvant study. Breast Cancer Res 17:32. https://doi.org/10.1186/s13058-0150540-0 28. Du J, Liu J, Smith BJ, Tsao MS, Cullen JJ (2011) Role of Rac1-dependent NADPH oxidase in the growth of pancreatic cancer. Cancer Gene Ther 18(2):135–143. https://doi.org/10.1038/ cgt.2010.64. Epub 2010 Oct 29 29. Duncan LJ, Seaton DA (1962) The treatment of diabetes mellitus with metformin. Br J Clin Pract 16:129–132 30. Everhart J, Wright D (1995) Diabetes mellitus as a risk factor for pancreatic cancer. A metaanalysis. JAMA 273(20):1605–1609

5

Role of Mitochondria in Pancreatic Metabolism, Diabetes, and Cancer

89

31. Fan Y, Gan Y, Shen Y, Cai X, Song Y, Zhao F, Yao M, Gu J, Tu H (2015) Leptin signaling enhances cell invasion and promotes the metastasis of human pancreatic cancer via increasing MMP-13 production. Oncotarget 6(18):16120–16134 32. Frantz C, Karydis A, Nalbant P, Hahn KM, Barber DL (2007) Positive feedback between Cdc42 activity and H+ efflux by the Na-H exchanger NHE1 for polarity of migrating cells. J Cell Biol 179(3):403–410 33. Gaglio D, Metallo CM, Gameiro PA, Hiller K, Danna LS, Balestrieri C, Alberghina L, Stephanopoulos G, Chiaradonna F (2011) Oncogenic K-Ras decouples glucose and glutamine metabolism to support cancer cell growth. Mol Syst Biol 7:523. https://doi.org/10.1038/msb. 2011.56 34. Gapstur SM, Gann PH, Lowe W, Liu K, Colangelo L, Dyer A (2000) Abnormal glucose metabolism and pancreatic cancer mortality. JAMA 283(19):2552–2558 35. Gasparre G, Porcelli AM, Lenaz G, Romeo G (2013) Relevance of mitochondrial genetics and metabolism in cancer development. Cold Spring Harb Perspect Biol 5(2):pii: a011411. https:// doi.org/10.1101/cshperspect.a011411 36. Gebregiworgis T, Purohit V, Shukla SK, Tadros S, Chaika NV, Abrego J, Mulder SE, Gunda V, Singh PK (2017) Powers R1 glucose limitation alters glutamine metabolism in MUC1-overexpressing pancreatic Cancer cells. J Proteome Res 16(10):3536–3546. https:// doi.org/10.1021/acs.jproteome.7b00246. Epub 2017 Aug 30 37. Gross MI, Demo SD, Dennison JB, Chen L, Chernov-Rogan T, Goyal B, Janes JR, Laidig GJ, Lewis ER, Li J, Mackinnon AL, Parlati F, Rodriguez ML, Shwonek PJ, Sjogren EB, Stanton TF, Wang T, Yang J, Zhao F (2014) Bennett MK antitumor activity of the glutaminase inhibitor CB-839 in triple-negative breast cancer. Mol Cancer Ther 13(4):890–901. https:// doi.org/10.1158/1535-7163.MCT-13-0870. Epub 2014 Feb 12 38. Guillaumond F, Leca J, Olivares O, Lavaut MN, Vidal N, Berthezène P, Dusetti NJ, Loncle C, Calvo E, Turrini O, Iovanna JL, Tomasini R, Vasseur S (2013) Strengthened glycolysis under hypoxia supports tumor symbiosis and hexosamine biosynthesis in pancreatic adenocarcinoma. Proc Natl Acad Sci USA 110(10):3919–3924. https://doi.org/10.1073/pnas. 1219555110. Epub 2013 Feb 13 39. Haataja L, Gurlo T, Huang CJ, Butler PC (2008) Islet amyloid in type 2 diabetes, and the toxic oligomer hypothesis. Endocr Rev 29(3):303–316. https://doi.org/10.1210/er.2007-0037. Epub 2008 Feb 26 40. Halbrook CJ, Lyssiotis CA (2017) Employing metabolism to improve the diagnosis and treatment of pancreatic cancer. Cancer Cell 31(1):5–19. https://doi.org/10.1016/j.ccell.2016. 12.006 41. Hamasaki M, Furuta N, Matsuda A, Nezu A, Yamamoto A, Fujita N, Oomori H, Noda T, Haraguchi T, Hiraoka Y, Amano A, Yoshimori T (2013) Autophagosomes form at ERmitochondria contact sites. Nature 495(7441):389–393. https://doi.org/10.1038/nature11910. Epub 2013 Mar 3 42. Hanahan D, Weinberg RA (2011) Hallmarks of cancer: the next generation. Cell 144 (5):646–674. https://doi.org/10.1016/j.cell.2011.02.013 43. Hanahan D, Weinberg RA (2000) The hallmarks of cancer. Cell 100(1):57–70 44. Hernandez MG, Aguilar AG, Burillo J, Oca RG, Manca MA, Novials A, Alcarraz-Vizan G (2018) Guillén C3,4, Benito M1,5. Pancreatic β cells overexpressing hIAPP impaired mitophagy and unbalanced mitochondrial dynamics. Cell Death Dis 9(5):481. https://doi. org/10.1038/s41419-018-0533-x 45. Holmstrom KM, Finkel T (2014) Cellular mechanisms and physiological consequences of redox-dependent signalling. Nat Rev Mol Cell Biol 15(6):411–421. https://doi.org/10.1038/ nrm3801 46. Hoshino A, Ariyoshi M, Okawa Y, Kaimoto S, Uchihashi M, Fukai K, Iwai-Kanai E, Ikeda K, Ueyama T, Ogata T, Matoba S (2014) Inhibition of p53 preserves Parkin-mediated mitophagy and pancreatic β-cell function in diabetes. Proc Natl Acad Sci USA 111(8):3116–3121. https:// doi.org/10.1073/pnas.1318951111. Epub 2014 Feb 10

90

N. K. Talari et al.

47. Hu Y, Lu W, Chen G, Wang P, Chen Z, Zhou Y, Ogasawara M, Trachootham D, Feng L, Pelicano H, Chiao PJ, Keating MJ, Garcia-Manero G, Huang P (2012) K-ras(G12V) transformation leads to mitochondrial dysfunction and a metabolic switch from oxidative phosphorylation to glycolysis. Cell Res 22(2):399–412. https://doi.org/10.1038/cr.2011.145. Epub 2011 Aug 30 48. Hull RL, Westermark GT, Westermark P, Kahn SE (2004) Islet amyloid: a critical entity in the pathogenesis of type 2 diabetes. J Clin Endocrinol Metab 89(8):3629–3643 49. Huxley R, Ansary-Moghaddam A, Berrington de González A, Barzi F, Woodward M (2005) Type-II diabetes and pancreatic cancer: a meta-analysis of 36 studies. Br J Cancer 92 (11):2076–2083 50. Jacobetz MA, Chan DS, Neesse A, Bapiro TE, Cook N, Frese KK, Feig C, Nakagawa T, Caldwell ME, Zecchini HI, Lolkema MP, Jiang P, Kultti A, Thompson CB, Maneval DC, Jodrell DI, Frost GI, Shepard HM, Skepper JN, Tuveson DA (2013) Hyaluronan impairs vascular function and drug delivery in a mouse model of pancreatic cancer. Gut 62(1):112–120. https://doi.org/10.1136/gutjnl-2012-302529. Epub 2012 Mar 30 51. Jeong SM, Hwang S, Seong RH (2016) Transferrin receptor regulates pancreatic cancer growth by modulating mitochondrial respiration and ROS generation. Biochem Biophys Res Commun 471(3):373–379. https://doi.org/10.1016/j.bbrc.2016.02.023. Epub 2016 Feb 8 52. Jia ZY, Shen TY, Jiang W, Qin HL (2017) Identification of molecular mechanisms of glutamine in pancreatic cancer. Oncol Lett 14(6):6395–6402. https://doi.org/10.3892/ol. 2017.7068. Epub 2017 Sep 26 53. Kalender A, Selvaraj A, Kim SY, Gulati P, Brûlé S, Viollet B, Kemp BE, Bardeesy N, Dennis P, Schlager JJ, Marette A, Kozma SC, Thomas G (2010) Metformin, independent of AMPK, inhibits mTORC1 in a rag GTPase-dependent manner. Cell Metab 11(5):390–401. https://doi.org/10.1016/j.cmet.2010.03.014 54. Karnoub AE, Weinberg RA (2008) Ras oncogenes: split personalities. Nat Rev Mol Cell Biol 9(7):517–531. https://doi.org/10.1038/nrm2438 55. Khaled AR, Kim K, Hofmeister R, Muegge K, Durum SK (1999) Withdrawal of IL-7 induces Bax translocation from cytosol to mitochondria through a rise in intracellular pH. Proc Natl Acad Sci USA 96(25):14476–14481 56. Kisfalvi K, Moro A, Sinnett-Smith J, Eibl G, Rozengurt E (2013) Metformin inhibits the growth of human pancreatic cancer xenografts. Pancreas 42(5):781–785. https://doi.org/10. 1097/MPA.0b013e31827aec40 57. Krishnan KJ, Greaves LC, Reeve AK, Turnbull DM (2007) Mitochondrial DNA mutations and aging. Ann N Y Acad Sci 1100:227–240 58. Kwong SC, Brubacher J (1998) Phenformin and lactic acidosis: a case report and review. J Emerg Med 16(6):881–886 59. Lau ST, Lin ZX, Leung PS (2010) Role of reactive oxygen species in brucein D-mediated p38mitogen-activated protein kinase and nuclear factor-kappaB signalling pathways in human pancreatic adenocarcinoma cells. Br J Cancer 102(3):583–593. https://doi.org/10.1038/sj.bjc. 6605487. Epub 2010 Jan 12 60. Lee JK, Edderkaoui M, Truong P, Ohno I, Jang KT, Berti A, Pandol SJ (2007) Gukovskaya AS NADPH oxidase promotes pancreatic cancer cell survival via inhibiting JAK2 dephosphorylation by tyrosine phosphatases. Gastroenterology 133(5):1637–1648. Epub 2007 Aug 14 61. Levine AJ, Puzio-Kuter AM (2010) The control of the metabolic switch in cancers by oncogenes and tumor suppressor genes. Science 330(6009):1340–1344. https://doi.org/10. 1126/science.1193494 62. Liang C, Qin Y, Zhang B, Ji S, Shi S, Xu W, Liu J, Xiang J, Liang D, Hu Q, Liu L, Liu C, Luo G, Ni Q, Xu J, Yu X (2016) Energy sources identify metabolic phenotypes in pancreatic cancer. Acta Biochim Biophys Sin Shanghai 48(11):969–979. Epub 2016 Sep 20 63. Liao C, Hu B, Arno MJ, Panaretou B (2007) Genomic screening in vivo reveals the role played by vacuolar H+ ATPase and cytosolic acidification in sensitivity to DNA-damaging agents such as cisplatin. Mol Pharmacol 71(2):416–425. Epub 2006 Nov 8

5

Role of Mitochondria in Pancreatic Metabolism, Diabetes, and Cancer

91

64. Liou GY, Döppler H, DelGiorno KE, Zhang L, Leitges M, Crawford HC, Murphy MP, Storz P (2016) Mutant KRas-induced mitochondrial oxidative stress in acinar cells upregulates EGFR signaling to drive formation of pancreatic precancerous lesions. Cell Rep 14(10):2325–2336. https://doi.org/10.1016/j.celrep.2016.02.029. Epub 2016 Mar 3 65. Lu H, Koshkin V, Allister EM, Gyulkhandanyan AV, Wheeler MB (2010) Molecular and metabolic evidence for mitochondrial defects associated with beta-cell dysfunction in a mouse model of type 2 diabetes. Diabetes 59(2):448–459. https://doi.org/10.2337/db09-0129. Epub 2009 Nov 10 66. Lu W, Hu Y, Chen G, Chen Z, Zhang H, Wang F, Feng L, Pelicano H, Wang H, Keating MJ, Liu J, McKeehan W, Wang H, Luo Y, Huang P (2012) Novel role of NOX in supporting aerobic glycolysis in cancer cells with mitochondrial dysfunction and as a potential target for cancer therapy. PLoS Biol 10(5):e1001326. https://doi.org/10.1371/journal.pbio.1001326. Epub 2012 May 8 67. Luo T, Masson K, Jaffe JD, Silkworth W, Ross NT, Scherer CA, Scholl C, Fröhling S, Carr SA, Stern AM, Schreiber SL, Golub TR (2012) STK33 kinase inhibitor BRD-8899 has no effect on KRAS-dependent cancer cell viability. Proc Natl Acad Sci USA 109(8):2860–2865. https://doi.org/10.1073/pnas.1120589109. Epub 2012 Feb 9 68. Lycan TW, Pardee TS, Petty WJ, Bonomi M, Alistar A, Lamar ZS, Isom S, Chan MD, Miller AA, Ruiz J (2016) A phase II clinical trial of CPI-613 in patients with relapsed or refractory small cell lung carcinoma. PLoS One 11(10):e0164244. https://doi.org/10.1371/journal.pone. 0164244. eCollection 2016 69. Ma J, Sawai H, Matsuo Y, Ochi N, Yasuda A, Takahashi H, Wakasugi T, Funahashi H, Sato M, Takeyama H (2010) IGF-1 mediates PTEN suppression and enhances cell invasion and proliferation via activation of the IGF-1/PI3K/Akt signaling pathway in pancreatic cancer cells. J Surg Res 160(1):90–101. https://doi.org/10.1016/j.jss.2008.08.016. Epub 2008 Sep 13 70. Madiraju AK, Erion DM, Rahimi Y, Zhang XM, Braddock DT, Albright RA, Prigaro BJ, Wood JL, Bhanot S, MacDonald MJ, Jurczak MJ, Camporez JP, Lee HY, Cline GW (2014) Samuel VT2, Kibbey RG8, Shulman GI9. Metformin suppresses gluconeogenesis by inhibiting mitochondrial glycerophosphate dehydrogenase. Nature 510(7506):542–546. https://doi.org/10.1038/nature13270. Epub 2014 May 21 71. Maisonneuve P, Lowenfels AB (2015) Risk factors for pancreatic cancer: a summary review of meta-analytical studies. Risk factors for pancreatic cancer: a summary review of metaanalytical studies. Int J Epidemiol 44(1):186–198. https://doi.org/10.1093/ije/dyu240. Epub 2014 Dec 14 72. Marshall S (2006) Role of insulin, adipocyte hormones, and nutrient-sensing pathways in regulating fuel metabolism and energy homeostasis: a nutritional perspective of diabetes, obesity, and cancer. Sci STKE 2006(346):re7 73. Maurer T, Garrenton LS, Oh A, Pitts K, Anderson DJ, Skelton NJ, Fauber BP, Pan B, Malek S, Stokoe D, Ludlam MJ, Bowman KK, Wu J, Giannetti AM, Starovasnik MA, Mellman I, Jackson PK, Rudolph J, Wang W, Fang G (2012) Small-molecule ligands bind to a distinct pocket in Ras and inhibit SOS-mediated nucleotide exchange activity. Proc Natl Acad Sci USA 109(14):5299–5304. https://doi.org/10.1073/pnas.1116510109. Epub 2012 Mar 19 74. Michalski CW, Hackert T, Büchler MW (2017) Targeting metabolism in pancreatic cancer. Lancet Oncol 18(6):699–700. https://doi.org/10.1016/S1470-2045(17)30304-2. Epub 2017 May 8 75. Mochizuki T, Furuta S, Mitsushita J, Shang WH, Ito M, Yokoo Y, Yamaura M, Ishizone S, Nakayama J, Konagai A, Hirose K, Kiyosawa K, Kamata T (2006) Inhibition of NADPH oxidase 4 activates apoptosis via the AKT/apoptosis signal-regulating kinase 1 pathway in pancreatic cancer PANC-1 cells. Oncogene 25(26):3699–3707. Epub 2006 Mar 13 76. Moreno-Sanchez R, Marín-Hernández A, Saavedra E, Pardo JP, Ralph SJ, RodríguezEnríquez S (2014) Who controls the ATP supply in cancer cells? Biochemistry lessons to understand cancer energy metabolism. Int J Biochem Cell Biol 50:10–23. https://doi.org/10. 1016/j.biocel.2014.01.025. Epub 2014 Feb 7 77. Ni HM, Williams JA, Ding WX (2015) Mitochondrial dynamics and mitochondrial quality control. Redox Biol 4:6–13. https://doi.org/10.1016/j.redox.2014.11.006. Epub 2014 Nov 20

92

N. K. Talari et al.

78. Ozcan U, Yilmaz E, Ozcan L, Furuhashi M, Vaillancourt E, Smith RO, Görgün CZ, Hotamisligil GS (2006) Chemical chaperones reduce ER stress and restore glucose homeostasis in a mouse model of type 2 diabetes. Science 313(5790):1137–1140 79. Papke B, Murarka S, Vogel HA, Martín-Gago P, Kovacevic M, Truxius DC, Fansa EK, Ismail S, Zimmermann G, Heinelt K, Schultz-Fademrecht C, Al Saabi A, Baumann M, Nussbaumer P, Wittinghofer A, Waldmann H, Bastiaens P (2016) Identification of pyrazolopyridazinones as PDEδ inhibitors. Nat Commun 7:11360. https://doi.org/10.1038/ncomms11360 80. Pardee TS, Lee K, Luddy J, Maturo C, Rodriguez R, Isom S, Miller LD, Stadelman KM, Levitan D, Hurd D, Ellis LR, Harrelson R, Manuel M, Dralle S, Lyerly S, Powell BL (2014) A phase I study of the first-in-class antimitochondrial metabolism agent, CPI-613, in patients with advanced hematologic malignancies. Clin Cancer Res 20(20):5255–5264. https://doi.org/ 10.1158/1078-0432.CCR-14-1019. Epub 2014 Aug 27 81. Park MT, Kim MJ, Suh Y, Kim RK, Kim H, Lim EJ, Yoo KC, Lee GH, Kim YH, Hwang SG, Yi JM, Lee SJ (2014) Novel signaling axis for ROS generation during K-Ras-induced cellular transformation. Cell Death Differ 21(8):1185–1197. https://doi.org/10.1038/cdd.2014.34. Epub 2014 Mar 14 82. Parks SK, Chiche J, Pouysségur J (2013) Disrupting proton dynamics and energy metabolism for cancer therapy. Nat Rev Cancer 13(9):611–623. https://doi.org/10.1038/nrc3579 83. Permert J, Larsson J, Westermark GT, Herrington MK, Christmanson L, Pour PM et al (1994) Islet amyloid polypeptide in patients with pancreatic cancer and diabetes. N Engl J Med 330:313–318. https://doi.org/10.1056/NEJM199402033300503 84. Pollak MN (2012) Investigating metformin for cancer prevention and treatment: the end of the beginning. Cancer Discov 2(9):778–790. https://doi.org/10.1158/2159-8290.CD-12-0263. Epub 2012 Aug 27 85. Pollak MN, Schernhammer ES, Hankinson SE (2004) Insulin-like growth factors and neoplasia. Nat Rev Cancer 4(7):505–518 86. Powell DR, Suwanichkul A, Cubbage ML, DePaolis LA, Snuggs MB, Lee PD (1991) Insulin inhibits transcription of the human gene for insulin-like growth factor-binding protein-1. J Biol Chem 266(28):18868–18876 87. Provenzano PP, Cuevas C, Chang AE, Goel VK, Von Hoff DD, Putney LK, Barber DL (2003) Na-H exchange-dependent increase in intracellular pH times G2/M entry and transition. J Biol Chem 278(45):44645–44649. Epub 2003 Aug 28 88. Putney LK, Barber DL (2003) Na-H exchange-dependent increase in intracellular pH times G2/M entry and transition. J Biol Chem Nov 7;278(45):44645–9. Epub 2003 Aug 28 89. Qu Y, Sun L, Yang Z, Han R (2011) Variation in the PTEN-induced putative kinase 1 gene associated with the increase risk of type 2 diabetes in northern Chinese. J Genet 90(1):125–128 90. Racker E, Resnick RJ, Feldman R (1985) Glycolysis and methylaminoisobutyrate uptake in rat-1 cells transfected with ras or myc oncogenes. Proc Natl Acad Sci USA 82(11):3535–3538 91. Rajeshkumar NV, Dutta P, Yabuuchi S, de Wilde RF, Martinez GV, Le A, Kamphorst JJ, Rabinowitz JD, Jain SK, Hidalgo M, Dang CV, Gillies RJ, Maitra A (2015) Therapeutic targeting of the Warburg effect in pancreatic cancer relies on an absence of p53 function. Cancer Res 75(16):3355–3364. https://doi.org/10.1158/0008-5472.CAN-15-0108. Epub 2015 Jun 25 92. Rozengurt E, Sinnett-Smith J, Kisfalvi K (2010) Crosstalk between insulin/insulin-like growth factor-1 receptors and G protein-coupled receptor signaling systems: a novel target for the antidiabetic drug metformin in pancreatic cancer. Clin Cancer Res 16(9):2505–2511. https:// doi.org/10.1158/1078-0432.CCR-09-2229. Epub 2010 Apr 13 93. Sancak Y, Peterson TR, Shaul YD, Lindquist RA, Thoreen CC, Bar-Peled L, Sabatini DM (2008) The Rag GTPases bind raptor and mediate amino acid signaling to mTORC1. Science 320(5882):1496–1501. https://doi.org/10.1126/science.1157535. Epub 2008 May 22 94. Savai R, Pullamsetti SS, Banat GA, Weissmann N, Ghofrani HA, Grimminger F, Schermuly RT (2010) Targeting cancer with phosphodiesterase inhibitors. Expert Opin Investig Drugs 19 (1):117–131. https://doi.org/10.1517/13543780903485642 95. Scheele C, Nielsen AR, Walden TB, Sewell DA, Fischer CP, Brogan RJ, Petrovic N, Larsson O, Tesch PA, Wennmalm K, Hutchinson DS, Cannon B, Wahlestedt C, Pedersen BK,

5

Role of Mitochondria in Pancreatic Metabolism, Diabetes, and Cancer

93

Timmons JA (2007) Altered regulation of the PINK1 locus: a link between type 2 diabetes and neurodegeneration? FASEB J 21(13):3653–3665. Epub 2007 Jun 12 96. Segal ED, Yasmeen A, Beauchamp MC, Rosenblatt J, Pollak M, Gotlieb WH (2011) Relevance of the OCT1 transporter to the antineoplastic effect of biguanides. Biochem Biophys Res Commun 414(4):694–699. https://doi.org/10.1016/j.bbrc.2011.09.134. Epub 2011 Oct 2 97. Shelton LM, Huysentruyt LC, Seyfried TN (2010) Glutamine targeting inhibits systemic metastasis in the VM-M3 murine tumor model. Int J Cancer 127(10):2478–2485. https://doi. org/10.1002/ijc.25431 98. Shima F, Yoshikawa Y, Ye M, Araki M, Matsumoto S, Liao J, Hu L, Sugimoto T, Ijiri Y, Takeda A, Nishiyama Y, Sato C, Muraoka S, Tamura A, Osoda T, Tsuda K, Miyakawa T, Fukunishi H, Shimada J, Kumasaka T, Yamamoto M, Kataoka T (2013) In silico discovery of small-molecule Ras inhibitors that display antitumor activity by blocking the Ras-effector interaction. Proc Natl Acad Sci USA 110(20):8182–8187. https://doi.org/10.1073/pnas. 1217730110. Epub 2013 Apr 29 99. Sinnett-Smith J, Kisfalvi K, Kui R, Rozengurt E (2013) Metformin inhibition of mTORC1 activation, DNA synthesis and proliferation in pancreatic cancer cells: dependence on glucose concentration and role of AMPK. Biochem Biophys Res Commun 430(1):352–357. https:// doi.org/10.1016/j.bbrc.2012.11.010. Epub 2012 Nov 15 100. Smith GD, Egger M, Shipley MJ, Marmot MG (1992) Post-challenge glucose concentration, impaired glucose tolerance, diabetes, and cancer mortality in men. Am J Epidemiol 136 (9):1110–1114 101. Son J, Lyssiotis CA, Ying H, Wang X, Hua S, Ligorio M, Perera RM, Ferrone CR, Mullarky E, Shyh-Chang N, Kang Y, Fleming JB, Bardeesy N, Asara JM, Haigis MC, DePinho RA, Cantley LC, Kimmelman AC (2013) Glutamine supports pancreatic cancer growth through a KRAS-regulated metabolic pathway. Nature 496(7443):101–105. https://doi.org/10.1038/ nature12040. Epub 2013 Mar 27 102. Sousa CM, Biancur DE, Wang X, Halbrook CJ, Sherman MH, Zhang L, Kremer D, Hwang RF, Witkiewicz AK, Ying H, Asara JM, Evans RM, Cantley LC, Lyssiotis CA, Kimmelman AC (2016) Pancreatic stellate cells support tumour metabolism through autophagic alanine secretion. Nature 536(7617):479–483. https://doi.org/10.1038/nature19084. Epub 2016 Aug 10 103. Stolzenberg-Solomon RZ, Graubard BI, Chari S, Limburg P, Taylor PR, Virtamo J, Albanes D (2005) Insulin, glucose, insulin resistance, and pancreatic cancer in male smokers. JAMA 294 (22):2872–2878 104. Stolzenberg-Solomon RZ, Newton CC, Silverman DT, Pollak M, Nogueira LM, Weinstein SJ, Albanes D, Männistö S, Jacobs EJ (2015) Circulating leptin and risk of pancreatic Cancer: a pooled analysis from 3 cohorts. Am J Epidemiol 182(3):187–197. https://doi.org/10.1093/aje/ kwv041. Epub 2015 Jun 17 105. Sun Q, Burke JP, Phan J, Burns MC, Olejniczak ET, Waterson AG, Lee T, Rossanese OW, Fesik SW (2012) Discovery of small molecules that bind to K-Ras and inhibit Sos-mediated activation. Angew Chem Int Ed Engl 51(25):6140–6143. https://doi.org/10.1002/anie. 201201358. Epub 2012 May 8 106. Szabadkai G, Duchen MR (2009) Mitochondria mediated cell death in diabetes. Apoptosis 14 (12):1405–1423. https://doi.org/10.1007/s10495-009-0363-5 107. Teoh ML, Sun W, Smith BJ, Oberley LW, Cullen JJ (2007) Modulation of reactive oxygen species in pancreatic cancer. Clin Cancer Res 13(24):7441–7450 108. Vaquero EC, Edderkaoui M, Pandol SJ, Gukovsky I, Gukovskaya AS (2004) Reactive oxygen species produced by NAD(P)H oxidase inhibit apoptosis in pancreatic cancer cells. J Biol Chem 279(33):34643–34654. Epub 2004 May 23 109. Wallace DC (2012) Mitochondria and cancer. Nat Rev Cancer 12(10):685–698. https://doi. org/10.1038/nrc3365 110. Wang JB, Erickson JW, Fuji R, Ramachandran S, Gao P, Dinavahi R, Wilson KF, Ambrosio AL, Dias SM, Dang CV, Cerione RA (2010) Targeting mitochondrial glutaminase activity inhibits oncogenic transformation. Cancer Cell 18(3):207–219. https://doi.org/10.1016/j.ccr. 2010.08.009

94

N. K. Talari et al.

111. Wang P, Song M, Zeng ZL, Zhu CF, Lu WH, Yang J, Ma MZ, Huang AM, Hu Y, Huang P (2015) Identification of NDUFAF1 in mediating K-Ras induced mitochondrial dysfunction by a proteomic screening approach. Oncotarget 6(6):3947–3962 112. Webb BA, Chimenti M, Jacobson MP, Barber DL (2011) Dysregulated pH: a perfect storm for cancer progression. Nat Rev Cancer 11(9):671–677. https://doi.org/10.1038/nrc3110 113. Wei EK, Giovannucci E, Fuchs CS, Willett WC, Mantzoros CS (2005) Low plasma adiponectin levels and risk of colorectal cancer in men: a prospective study. J Natl Cancer Inst 97(22):1688–1694 114. Weinberg F, Hamanaka R, Wheaton WW, Weinberg S, Joseph J, Lopez M, Kalyanaraman B, Mutlu GM, Budinger GR, Chandel NS (2010) Mitochondrial metabolism and ROS generation are essential for Kras-mediated tumorigenicity. Proc Natl Acad Sci USA 107(19):8788–8793. https://doi.org/10.1073/pnas.1003428107. Epub 2010 Apr 26 115. Weydert C, Roling B, Liu J, Hinkhouse MM, Ritchie JM, Oberley LW, Cullen JJ (2003) Suppression of the malignant phenotype in human pancreatic cancer cells by the overexpression of manganese superoxide dismutase. Mol Cancer Ther 2(4):361–369 116. Wu Y, Lu J, Antony S, Juhasz A, Liu H, Jiang G, Meitzler JL, Hollingshead M, Haines DC, Butcher D, Roy K, Doroshow JH (2013) Activation of TLR4 is required for the synergistic induction of dual oxidase 2 and dual oxidase A2 by IFN-γ and lipopolysaccharide in human pancreatic cancer cell lines. J Immunol 190(4):1859–1872. https://doi.org/10.4049/jimmunol. 1201725. Epub 2013 Jan 7 117. Wullschleger S, Loewith R, Hall MN (2006) TOR signaling in growth and metabolism. Cell 124(3):471–484 118. Xiang Y, Stine ZE, Xia J, Lu Y, O'Connor RS, Altman BJ, Hsieh AL, Gouw AM, Thomas AG, Gao P, Sun L, Song L, Yan B, Slusher BS, Zhuo J, Ooi LL, Lee CG, Mancuso A, McCallion AS, Le A, Milone MC, Rayport S, Felsher DW, Dang CV (2015) Targeted inhibition of tumorspecific glutaminase diminishes cell-autonomous tumorigenesis. J Clin Invest 125(6):2293– 2306. https://doi.org/10.1172/JCI75836. Epub 2015 Apr 27 119. Yang S, Wang X, Contino G, Liesa M, Sahin E, Ying H, Bause A, Li Y, Stommel JM, Dell'antonio G, Mautner J, Tonon G, Haigis M, Shirihai OS, Doglioni C, Bardeesy N, Kimmelman AC (2011) Pancreatic cancers require autophagy for tumor growth. Genes Dev 25(7):717–729. https://doi.org/10.1101/gad.2016111. Epub 2011 Mar 15 120. Yeo TP (2015) Demographics, epidemiology, and inheritance of pancreatic ductal adenocarcinoma. Semin Oncol 42(1):8–18. https://doi.org/10.1053/j.seminoncol.2014.12.002. Epub 2014 Dec 9 121. Ying H, Kimmelman AC, Lyssiotis CA, Hua S, Chu GC, Fletcher-Sananikone E, Locasale JW, Son J, Zhang H, Coloff JL, Yan H, Wang W, Chen S, Viale A, Zheng H, Paik JH, Lim C, Guimaraes AR, Martin ES, Chang J, Hezel AF, Perry SR, Hu J, Gan B, Xiao Y, Asara JM, Weissleder R, Wang YA, Chin L, Cantley LC, DePinho RA (2012) Oncogenic Kras maintains pancreatic tumors through regulation of anabolic glucose metabolism. Cell 149(3):656–670. https://doi.org/10.1016/j.cell.2012.01.058 122. Zhang CY, Baffy G, Perret P, Krauss S, Peroni O, Grujic D, Hagen T, Vidal-Puig AJ, Boss O, Kim YB, Zheng XX, Wheeler MB, Shulman GI, Chan CB, Lowell BB (2001) Uncoupling protein-2 negatively regulates insulin secretion and is a major link between obesity, beta cell dysfunction, and type 2 diabetes. Cell 105(6):745–755 123. Zhao D, Zou SW, Liu Y, Zhou X, Mo Y, Wang P, Xu YH, Dong B, Xiong Y, Lei QY, Guan KL (2013) Lysine-5 acetylation negatively regulates lactate dehydrogenase A and is decreased in pancreatic cancer. Cancer Cell 23(4):464–476. https://doi.org/10.1016/j.ccr.2013.02.005. Epub 2013 Mar 21 124. Zyromski NJ, Mathur A, Pitt HA, Wade TE, Wang S, Nakshatri P, Swartz-Basile DA, Nakshatri H (2009) Obesity potentiates the growth and dissemination of pancreatic cancer. Surgery 146(2):258–263. https://doi.org/10.1016/j.surg.2009.02.024

6

Targeting Mitochondrial Enzymes in Pancreatic Cancer Gowru Srivani, Begum Dariya, Afroz Alam, and Ganji Purnachandra Nagaraju

Abstract

Mitochondria is a well-known dynamic intracellular organelle. It performs emerging functions in the cell, such as production of energy (ATP generators), apoptosis, and cell cycle regulation. Over the past years, researchers have concentrated on the complexity of pancreatic cancer by mapping genetic mutations associated with it. In these efforts, the role of mitochondrion to the pathogenesis of pancreatic cancer has been neglected. However, recently a growing body of research showed the evidence that mitochondrial enzymes play a crucial role in development of pancreatic cancer. In fact, deregulated mitochondrial enzymes not only participate in the metabolic reprogramming of tumor cells but also modulate cellular processes involved in tumor proliferation and progression. In this review, we described the association between the deregulated mitochondrial enzymes and development of pancreatic cancer. Keywords

Mitochondria · ATP · Pancreatic cancer and mitochondrial enzymes

Abbreviations 2-HG 2SC AcCoA

2-hydroxyglutarate S-2-(succino)cysteine Acetyl coenzyme A

G. Srivani · B. Dariya · A. Alam Department of Bioscience and Biotechnology, Banasthali University, Vanasthali, Rajasthan, India G. P. Nagaraju (*) Department of Hematology and Medical Oncology, Winship Cancer Institute, Emory University, Atlanta, GA, USA e-mail: [email protected] # Springer Nature Singapore Pte Ltd. 2019 G. P. Nagaraju, A. BM Reddy (eds.), Exploring Pancreatic Metabolism and Malignancy, https://doi.org/10.1007/978-981-32-9393-9_6

95

96

Aco1 AMPK ATP Ca CAFs CS EMT ETC FADH FH FHD FMN HIF HLRCC ICDH IMM IRE IRP1 JHDMS KDMS KEAP MDH MTATP6 NADPH NRF2 OAA OG2 OMM OXPHOS PC PDAC PGC/PCC PHD PRMT4 R-2-HG ROS SDH Succ CoA TCA TET α-KG

G. Srivani et al.

Aconitase1 AMP-activated protein kinase Adenosine triphosphate Calcium Cancer-associated fibroblasts Citrate synthase Epithelial to mesenchymal transition Electron transport chain Flavin adenine dinucleotide Fumarate dehydrogenase Fumarate hydratase Flavin mononucleotides Hypoxia-inducible factor Hereditary leiomyomatosis and renal cell cancer Isocitrate dehydrogenase Inner membrane of mitochondrion Iron-responsive elements Iron regulatory protein-1 Jumonji-C histone demethylases Histone lysine demethylases Kelch-like ECH-associated protein-1 Malate dehydrogenase mt DNA gene encoding for the CV subunit 6 Nicotinamide adenine dinucleotide phosphate Nuclear factor erythroid 2-related factor-2 Oxaloacetate Oxoglutarate Outer membrane of mitochondrion Oxidative phosphorylation Pancreatic cancer Pancreatic ductal adenocarcinoma Paraganglioma and pheochromocytoma Prolyl hydroxylases Protein arginine methyltransferase 4 2-Hydroxyl glutamate Reactive oxygen species Succinate dehydrogenase Succinal coenzyme A Tricarboxylic acid Ten-eleven translocation family α-Ketoglutarate

6

Targeting Mitochondrial Enzymes in Pancreatic Cancer

6.1

97

Introduction

Pancreatic cancer (PC) is one of the fourth most aggressive cancers, diagnosed at a very late stage, with poor survival rate [1, 2]. It has greater metastatic potential and resistance to chemoradiotherapy [3]. Its development and resistance to treatment depend on the inhibition of the apoptosis. Although mitochondrion and their enzymes are identified as crucial regulators for cell death, their significance as target for cancer therapy has not been clinically explored [4]. Mainly, its initiation is associated with multiple interactions between host and causative factors. These interactions induce the genetic and metabolic alterations. The deregulated cellular energetics are correlated with mitochondrial dysfunction caused by defects in mitochondrial enzymes, altered tumor suppressor genes, as well as mutations (point mutation and copy number change) in the mitochondrial DNA. The mitochondrial enzymes including citrate synthase, succinate dehydrogenase (SDH), fumarate hydratase (FHD), and isocitrate dehydrogenase (ICDH) [5] are defected and are thus connected with sporadic or familial form of cancer. This contributes to the deregulating cellular energy, upregulation of enzymes, increasing proliferative signaling, activating metastasis, invasion, endorsing inflammation, angiogenesis, and resistance to apoptosis [6]. Mitochondria are an important master regulator of intracellular organelles present in the cell, which are involved in biosynthetic and bioenergetic functions to reach metabolic need of the cell [5, 7]. In addition, participated in the cellular homeostasis, like the formation of ATP via electron transport chain (ETC), and oxidative phosphorylation (OXPHOS) in combination with the oxidation of metabolic intermediates through TCA (tricarboxylic/citric acid) cycle. Moreover, it also participates in fatty acid catabolism by β-oxidation; production of heme, steroids, and pyrimidine; and production of excess reactive oxygen species (ROS) via regulation of calcium concentrations (Ca2+) and through trafficking of small metabolic substances. It is associated with the initiation and implementation of the cell senescence [8, 9]. Therefore, the abnormal communication between the mitochondrion and other cells may cause the changes in cellular homeostasis and organismal dysfunction. Certainly, dysfunction of mitochondrion has been associated with various pathological conditions such as neurodegenerative disorders, muscular degeneration, cardiovascular disorders, and cancer [10, 11]. Historically, the association between cancer cells and dysfunction of mitochondrion aimed at the metabolism [12]. In 1956, Warburg discovered the association between cancer cell, mitochondrial dysfunction [13], and the altered mitochondrial function that modulates the gene function, cell cycle, cell viability, and metabolism [14]. In this review, we discussed about the key role of altered mitochondrial enzymes and mitochondrial DNA mutations in initiating a complex cellular reprogramming that supports the development of PC growth.

98

6.2

G. Srivani et al.

Mitochondria and Their Enzymes

Mitochondria are an important organelle for cellular activities. It is the powerhouse of the cell. It is a double membrane organelle [15] and differs from organism to organism. The number of mitochondrion varies per tissue, organelle, and cell type. There are no mitochondrion present in red blood cells, and more than 2000 mitochondrion are present in the liver cells [16]. Mitochondria has a double-stranded circular DNA with 16.6 kb molecular weight, 22 transfer RNAs, 13 respiratory enzyme complex polypeptides, and 2 ribosomal RNAs that are involved in posttranslation of mitochondrial protein synthesis and respiration [17]. Mitochondria are differentiated from the cytoplasm through its outer and inner membranes. The outer membrane of mitochondrion (OMM) is characterized porous. The small uncharged molecules and ions can be freely crossed through the porins (pore-forming membrane proteins) like voltage-dependent anion channels (VDAC). Additionally, OMM also have particular translocases for importing large molecules such as protein molecules. Inner membrane of mitochondrion (IMM) is rigid and has a diffusion barrier for small molecules and ions. All these molecules or ions pass through the specific membrane transport proteins that are specific for particular molecule or ions. Thus, formed ion selectivity develops an electrochemical membrane potential ~ 180 mV in inner membrane. Further, the oxidative phosphorylation (OXPHOS) takes place in group of membrane protein complexes that generates the electrochemical gradient potential throughout the inner membrane and is utilized for the synthesis of ATP [18]. The OMM and IMM of mitochondrion are classified into three compartments, each having its significant role in protein synthesis. The innermost compartment is mitochondrial matrix, which is present surrounding the inner membrane. Mitochondrial matrix contains ribosomes, organic molecules, mitochondrial DNA, inorganic ions, nucleotide factors, and soluble enzymes. It is the place for DNA replication, transcription, synthesis of proteins, and various enzymatic reactions. The biosynthetic reactions of citric acid cycle, electron transport chain, respiration, and urea cycle occurred within the mitochondrial matrix. These cycles play a vital role to produce substrates for the formation of various macromolecules such as amino acids, nucleotides, heme, iron sulfur clusters, lipids, and nicotinamide adenine dinucleotide phosphate (NADPH) in regulated concentration for antioxidant protection [10]. The matrix contains various enzymes including citrate synthase, isocitrate dehydrogenase (IDH), alpha-ketoglutarate dehydrogenase, fumarate, succinate dehydrogenase (SDH), and malate dehydrogenase (MDH) that play a crucial role in citric acid cycle/ Krebs/TCA cycle and electron transport chain intermediates [10, 19]. These enzymes are frequently deregulated or mutated leading to tumorigenesis [20]. TCA cycle generates NADH and FADH2 through the electron transport chain that form a proton gradient throughout cristae (mitochondrial inner membrane) and produce ATP with the help of enzyme proton-ATP synthase. This permits protons to pass through the membrane in a unidirectional way, similar to chemiosmosis process [21, 22]. This metabolic process is called OXPHOS (oxidative phosphorylation), which oxidizes the nutrients to generate ATPs. In addition to

6

Targeting Mitochondrial Enzymes in Pancreatic Cancer

99

this, redox reactions also happen during this process that generates ROS (reactive oxygen species) such as peroxide or superoxide via oxygen reduction; ROS are harmful intermediates of OXPHOS. This process is of high potentiality as it reproduces 36 ATP molecules, produced during glycolysis. This process mainly takes place at complex-I, cytochrome-C oxidase, and complex-IV [23, 24]. The deregulated mitochondrial functions are mainly caused by dysfunction of mitochondrial enzymes or mutations in the mitochondrial DNA. These disturb the cellular bioenergetic process and support the metabolic reprogramming for tumor cell proliferation and survival as well as activating the tumor-promoting changes through the harmful intermediates like ROS, calcium, and small metabolites like glutamine produced by the mitochondrion. Moreover, various oncogenes or deregulated tumor suppressor proteins including P53, HIF (hypoxia-inducible factor), and RAS regulate cellular metabolism and mitochondrial respiration. These play a vital role in determining how the mitochondrial enzymes are utilized in the tumor cells. Thus, the altered tumor suppressors or oncogenes provide a direct association between the dysfunction of mitochondrion and tumorigenesis [19].

6.3

Dysfunction of Mitochondrial Enzymes and Pancreatic Cancer

Deregulated mitochondrial enzymes including citric acid cycle enzymes can alter the cellular energetics, resulting in the characteristic metabolic changes that are associated with the tumor progression and development. A series of biochemical reactions in the citric acid cycle and electron transport chain (ETC) catalyzed by various enzymes may play a vital role in pancreatic cancer initiation, proliferation, and progression (Fig. 6.1).

6.4

Citrate Synthase

Citrate synthase is the first and rate-limiting step in citric acid cycle. This is an irreversible condensation reaction of AcCoA (acetyl coenzyme A) and OAA (oxaloacetate) into citrate [25]. Continuously citrate from the matrix to the TCA cycle, this step is indeed crucial for transport of AcCoA in the form of citrate from mitochondrion to cytosol, which is then involved in synthesis of fatty acids or acetylation of proteins [26]. Overexpression and increased enzymatic activity of citrate synthase cause tumors, including pancreatic cancer, ovarian cancer, and renal oncocytoma [27]. Unfortunately, the changes made by citrate synthase like reprogramming the mitochondrial function have not been determined. Furthermore, there is no clarity how dysfunction of citrate synthase can promote the tumor proliferation and progression. Moreover, there are two conditions assumed – in one condition, the elevated levels of citrate synthase can produce excess amount of citrate, promoting tumor cells for the biosynthesis of fatty acids. This is mostly seen in the pancreatic cancer [28]. In the other condition, loss of citrate synthase

100

G. Srivani et al.

Fig. 6.1 Role of mitochondrial enzymes in pancreatic cancer development [85] Oncogenic metabolites R-2-HG, succinate, and fumarate are the substrates of the gain of function of cytosolic mitochondrial isoforms of IDH, SDH, and FH. These are intermediates of the citric acid cycle. Mutant tumor suppressor genes IDH, SDH, and FH generate the intracellular accumulation of succinate, fumarate, and 2-oxoglutarate (2-OG) dioxygenases, and also inhibit the PDH, KDMS, and TET, resulting in the HIF-1α gene modifications via epigenetic alterations that may be involved in the tumor development. Fumarate irreversibly alters the cysteine molecules in proteins through succination that leads to the activation of Keap1 and the integral activation of NRF2, resulting in the transcriptional genes associated with antioxidant response. Succination of the citric acid cycle enzyme aconitase 2 bound with the IRP1 proteins upregulated the HIF-1α activity and led to the unpaired aconitase function in FH-deficient cells IDH isocitrate dehydrogenase, SDH succinate dehydrogenase, FH fumarate dehydrogenase, MDH malate dehydrogenase, CS citrate synthase, PHD prolyl hydroxylases HIF hypoxia-inducible factor, TET ten-eleven translocation family, KDMS histone lysine demethylases, KEAP Kelch-like-ECH-associated protein-1, NRF2 nuclear factor erythroid 2-related factor-2, OAA oxalo acetate, AcCoA acetyl coenzyme A, RH R denotes HIF, DNA, histone, OG2 oxologltarate, R-2-HG 2-hydroxyl glutamate, Succ CoA succinal coenzyme A

could alter the mitochondrial dysfunction that activates glycolytic switch to promote tumor progression. In addition, loss of citrate synthase is also associated with the initiation of (EMT) epithelial to mesenchymal transition. This suggests that the loss of citrate synthase activity not only induces the metabolic reprogramming but also supports the tumor cell proliferation, invasion, and metastasis [29].

6.5

Aconitase

Aconitase is the second mitochondrial enzyme in the citric acid cycle. It converts citrate into isocitrate through the cis aconitase. It is also known as IRP1 (iron regulatory protein-1) and Fe-S cluster enzyme. It plays a major role in maintaining homeostasis of intracellular iron. Based upon the intracellular iron concentrations, it

6

Targeting Mitochondrial Enzymes in Pancreatic Cancer

101

may either act as an aconitase or exist attached to IRE (iron-responsive elements) in the untranslated regions of mRNA elements [30, 31]. Modified translated proteins control the uptake, excess utilization, and storage of iron [31]. In addition, these IRP1 play a vital role in iron absorption and erythropoiesis by regulating HIF-2α (hypoxia-inducible factor 2 α). Thus, these IRP1 or iron-regulating functions of aconitase are also associated with the tumor initiation and progression. The alteration in iron metabolism is indeed a key element for tumor cell survival [31]. In 2016, Yuan et al. [32] revealed that increased circulating levels of TCA cycle metabolites like aconitase and isocitrate and mutations in ACO1 (aconitase 1) gene are associated with tumor cell progression and metastasis. Moreover, this approximately increases twofold high risk for mortality of pancreatic cancer. Furthermore, their study data identified that ACO1 gene can be taken as the novel prognostic marker for pancreatic cancer.

6.6

Isocitrate Dehydrogenase

Isocitrate dehydrogenase (IDH) plays a vital role in cellular metabolism, tumor initiation, and metastasis. IDH catalyzes the conversion of isocitrate into 2-OG (2-oxoglutarate) with NADP as a cofactor. IDH exists in three isoforms: IDH1, IDH2, and IDH3. Among them, IDH3 is NADH (nicotinamide adenine dinucleotide) dependent, and the rest, IDH1 and IDH2, are NADPH (nicotinamide adenine dinucleotide phosphate) dependent [33, 34]. In the cytosol, IDH1 is encoded by the IDH1 gene which is located on 2q33.3. It is involved in oxidative decarboxylation of isocitrate to α-ketoglutarate to form NADPH from NADP+. The mitochondrion IDH2 is encoded by the IDH2 gene located on 15q26.1. It is associated with the reversible conversion of α-ketoglutarate to isocitrate and reduces NADPH to NAD+. Moreover, NADPH provides enough energy for reductive carboxylation [35]. Both IDH1 and IDH2 are homodimers that are functionally and structurally similar. They play a vital role in the oxidative reductase reactions and cellular activities against the oxidative destructions, where NADPH acts as a co factor. These reversible reductive decarboxylation also play a key role in various cellular processes including glycolysis and lipogenesis regulated through isocitrate transferase (ICT) that also produces citrate through the aconitase enzyme [36–38]. The NAD-dependent IDH3 plays a well-established function in citric acid cycle and, in addition, also catalyzes the irreversible conversion of isocitrate into α-ketoglutarate. NAD+ reduces NADH [34]. These reactions and isoform of IDH3 are mainly regulated by the substrate concentrations, negative and positive allosteric effectors like ATP/ADP ratio, calcium levels, citrate, and NAD+/NADH levels. Calcium, citrate, and ADP activate IDH3 function, whereas NADH, NADPH, and ATP suppress its function. Further, the α-ketoglutarate is metabolized into succinate which is a four-carbon molecule, and NADH is utilized in the ETC (electron transport chain) to produce ATP [39]. IDH3 is a heterotetramer having two α-subunits, one β-subunit, and one γ-subunit. IDH3 A gene is located on 15q25.1–

102

G. Srivani et al.

q25.2, IDH3 B on 20p13, and IDH3 G on Xq28 sequentially. Here, α-subunit acts as a catalytic unit, and the rest of the two β- and γ-subunits function as regulatory units [34, 40, 41]. Inherited IDH isoforms-cytoplasmic IDH1 and mitochondrial IDH2 have been identified in several cancers, including pancreatic cancer [42], colon cancer [43], glioma [44], acute myeloid leukemia [45], osteosarcoma [46], prostate cancer [47], glioblastoma [48], B-acute lymphoblastic leukemia [47], and intrahepatic cholangiocarcinoma [49]. The inherited IDH1 and IDH 2 convert α-KG (α-ketoglutarate) to reductive oncometabolite, i.e., 2-HG (2-hydroxyglutarate). The accumulation of high concentration of 2-HG in tumor cells enhances the cytokine independency, DNA, and histone methylation, which eventually results in the tumor cell dedifferentiation [50, 51]. The frequently occurring IDH mutants, IDH1-R132 and IDH2-R172, have the ability to produce 10–100-fold increased concentrations of R 2-HG. In 2012, Koivunen et al. showed in their study that mutant R-2-HGinduced PHD1 and PHD2 expression (also known as EgIN and EgIN2), thereby inhibiting HIF-1α activity, thus increases the tumor cell growth, proliferation, and cloning in soft-agar growth in astrocytes, indicating that HIF-1α in this regard may function as a tumor suppressor gene [52]. Recently in another study, Brody et al. revealed that IDH1 mutations were involved in the pancreatic cancer development and metastasis. This, for the first time, reveals a proof for intragenic mutations of IDH1 R132H in pancreatic cancer. As per the literature, IDH1 mutation-associated pancreatic cancer is very rare incident. Mutations in IDH1 causes the hypermethylation due to the increased level of 2-HG, which leads to neomorphic proliferation and dedifferentiation [53]. In support of this study, the identified mutant gain and the function of IDH1 have been found less efficient in catalyzing the reductive carboxylation reaction that converts α-KG (α-ketoglutarate) into isocitrate, leading to enhanced production of 2-HG. In addition, it was also observed that mutant IDH1 is responsible for the morphological changes and driver of epithelial to mesenchymal transition (EMT) in the later stages of pancreatic cancer, where it is extreamly susceptible to the harsh effects of hypoxia and chemotherapy [42, 54].

6.7

Succinate Dehydrogenase

Succinate dehydrogenase (SDH) is the only membrane-bound enzyme in the citric acid cycle. It is associated with the inner mitochondrial membrane that converts succinate (four-carbon molecule) into four-carbon molecule fumarate. It catalyzes the oxidation reaction without decarboxylation (reduction of FAD to FADH2) and produces FADH2. This also shows a unique connection between citric acid cycle and electron transport chain, where it is also notified as respiratory chain complex-II [54] and succinate-ubiquinone oxidoreductase. It is a highly conserved heterotetrameric enzyme complex that includes SDHA, SDHB, SDHC, and SDHD. SDHA and SDHB are mitochondrial matrix catalytic subunits; SDHA have FAD as a cofactor that binds to fumarate and succinate substrates; SDHB

6

Targeting Mitochondrial Enzymes in Pancreatic Cancer

103

contains the electron carriers of Fe-S center (iron-sulfur); SDHC and SDHD comprise the cytochrome b which is inner membrane protein and forms the complex-III (coenzyme Q-CoQ) binding site. The complex of SDH needs two factors: SDHAF1 (SDH assembly factor 1) and SDHAF2 [55, 56]. It is the only enzyme encoded by the nuclear DNA (nDNA) and lacks proton-pumping activity. Mutations in the SDH subunits are responsible for succinate accumulation. Thus, resulted metabolic reprogramming of the tumor microenvironment, despite normoxic conditions, provides a strategic atmosphere for tumor growth, proliferation, and survival. However, succinate is well known as a classical housekeeping gene [57]. Deactivating mutations of SDH subunits have been associated with various cancers like sporadic and hereditary cancers that include paraganglioma/pheochromocytoma (PGC/PCC) [58], breast cancer, pancreatic neuroendocrine tumor [59], renal cancer, and gastrointestinal stromal cancer [58]. Suppression of SDH enhances cytosolic, mitochondrial succinate levels and behaves as an oncogenic metabolite. The suppression of α-KG-PHDS (α-ketoglutarate-dependent prolyl hydroxylase) leads to stabilization of HIF-1α. Subsequently, it is translocated into the nucleus and induces the neovascularization and changes energy metabolism from oxidation to glycolysis [55, 60–62]. Stabilization of HIF-1α also resulted from the deactivation of PHDs via ROS which are produced from the respiratory complex-III; in addition, the lack of SDH cells enhanced the oxidative stress and ROS generation [63, 64]. Activation of HIF-1α under hypoxic conditions, together with succinate accumulation, enhances malignancy progression of pancreatic cancer [65], though SDH act as a tumor suppressor gene. Oncometabolites such as succinate and fumarate also suppress other α-KG-dependent dioxygenases, including TET family of 5-methylcytosine hydroxylases, histone, DNA demethylases, and JHDMs (Jumonji-C histone demethylases), resulting in epigenetic alterations and genome-wide changes of histone methylations [66].

6.8

Fumarate Hydratase

Another citric acid cycle enzyme is fumarate hydratase (FH), which catalyzes the oxidation reaction that converts fumarate into malate. Similar to SDH, FH also behaves as a tumor suppressor. However, germline-mutated FH is identified in HLRCC (hereditary leiomyomatosis and renal cell cancer) [67]. Recently, in another research, germline-mutated FH is also identified in pheochromocytoma and paraganglioma (PGC/PCC) [68] and is found downregulated in sporadic clear cell carcinoma and glioblastoma [69]. Fumarate possesses some common characteristics like 2-HG and succinate and has a unique feature associated with its chemical structure; it can suppress various OG-dependent enzymes, including histone, DNA demethylases, and PHDs [68]. Loss of FH function leads to the accumulation of intracellular fumarate, stabilization of HIF-1α, glycolysis, and activation of HIF-1-dependent pathways. Under physiological conditions, fumarate is indeed quietly reactive with α,

104

G. Srivani et al.

β-unsaturated electrophilic metabolites; it can covalently bind with cysteine residues in proteins to generate 2SC (S-(2-succino)-cysteine), and this process is called succination. This process stimulates loss of function in various proteins, including mitochondrial citric acid cycle enzyme aconitase, Keap1 (Kelch-like ECH-associated protein1), and activation of NRF2 (nuclear factor (erythroidderived2)-like protein) which is an antioxidant response sensor protein promoting the ROS signaling. Thus, the activation of various genes plays a vital role in tumor initiation and progression [70–73]. In one research study, it was observed that metabolic enzymes SDH and FH alterations of CAFs (cancer-associated fibroblasts) were associated with the development of pancreatic cancer with the upregulation of MiR-21. This study shows that FH also plays a key role in the progression of pancreatic cancer [74]. Similarly, another study suggested that FH mutations in kidney cancer are linked with decreased function of metabolic sensor and AMPK (AMP-activated protein kinase), resulting in the enhanced synthesis of proteins and fatty acid to support the continuing cellular anabolic process [75]. It is clear that FH is indirectly or directly linked with the development of various cancers, including pancreatic cancer.

6.9

Malate Dehydrogenase

Malate dehydrogenase is another key enzyme that reversibly converts malate into oxaloacetate using NAD/NADH coenzyme system [76]. There could be a special metabolic pathway in tumor mitochondrion that favors tumor progression, as several studies suggested that mitochondrial malate dehydrogenase can form binary complexes with citrate synthase, glutamate dehydrogenase, aspartate aminotransferase, and fumarase [77]. The upregulation of malic enzyme is observed in various types of human cancers. Furthermore, malic enzyme mRNA upregulation in bladder cancer tissue is observed and was suppressed due to overexpressed microRNA-612. It is clear that malate dehydrogenase is linked with cancer development [78]. Later, it was suggested that malate dehydrogenase 1(MDH1) is associated with pancreatic ductal adenocarcinoma (PDAC). Importantly, MDH1 R248 methylation was downregulated in human PDAC clinical samples and was observed along with low-level expression of protein arginine methyltransferase 4 (PRMT4/CARM1) [79]. In this regard, it is confirmed that MDH1 is a key metabolic enzyme in pancreatic and other cancers and could be a promising therapeutic target for drug development. Mutations in citric acid cycle enzymes CS, IDH, SDH, and FH have therefore been revealed as a significant metabolic reprogramming that also includes mitochondrial complex network, and thus, these modifications directly or indirectly support the development of tumorigenesis.

6

Targeting Mitochondrial Enzymes in Pancreatic Cancer

6.10

105

NADH-Ubiquinone Oxidoreductase

NADH-ubiquinone oxidoreductase is a well-known complex-I (CI) and is the largest electron transport complex, composed of NADH dehydrogenase, flavin mononucleotides (FMN), Fe-S center, and ubiquinone. It catalyzes the transfer of two electrons from the NADH to ubiquinone FMN and generates four protons and NAD+, which are further pumped from the mitochondrial matrix into intermembrane space for the generation of ATP [25]. It is the first place for receiving the electrons in the electron transport chain and is the active site for the generation of ROS. Hence, mutations in CI can expressively modify the redox homeostasis and bioenergetics [80]. Mutant mitochondrial genes encoding CI have been associated with the development of several malignancies including pancreas, breast, renal, thyroid, colon, prostate, and bladder as well medulloblastoma, head, and neck cancers [81].

6.11

Adenosine Triphosphate Synthase

Adenosine triphosphate (ATP) synthase is known as the complex V (CV), which is the last enzyme in the oxidative phosphorylation. CV generates ATP molecules from ADP and inorganic phosphate through the utilization of the electrochemical potential gradient across the inner mitochondrial space. It is recently found that ATP synthase is a part of PTP (permeability transition port) [82], which is a membrane-embedded mitochondrial complex. It is also involved in various mitochondrion-dependent events such as apoptosis and calcium buffering. Mutations occurred in ATP synthase encoded by the mtDNA have been involved in tumor development including pancreatic, thyroid, and prostate cancers [83]. Recently, Shidara et al. showed oncogenic functions of CV, and they observed two different point mutations in MTATP6 (mt DNA gene encoding for the CV subunit 6). Interestingly, mutant ATP 6 enhances the cell growth in 2D culture and showed increased oncogenic activity. Further, reintroducing the wild-type ATP6, which is nuclear encoded, results in inhibiting tumor growth. This study thus suggested mutant cells showing decreased activity of apoptosis enhanced ROS production and also clearly showed the linkage between CV mutations and tumor development [84] (Fig. 6.2).

6.12

Conclusion

In this review, we explained the link between mitochondrial enzymes and pancreatic cancer that are caused by mutations in the mitochondrial enzymes. We also discussed the dysfunction of mitochondrion and its importance in cancer development, genetic changes, and activities of oncogenic metabolites SDH, IDH, and FH involvement in progression of pancreatic cancer. Thus, these mitochondrial enzymes in pancreatic and other cancers could be a promising therapeutic target for drug development.

106

G. Srivani et al.

Fig. 6.2 Role of mutant IDH, SDH and, FH effect on ETC I and V resulting in the metabolic deregulation, modifying the amount of oxidative phosphorylation in tumor development. mt DNA mutations identified in NADH-dehydrogenase-UQ oxidoreductase (complex-I) and ATP synthase (complex-V) Complex-II succinate-UQ oxidoreductase, complex-III reduced UQ-Cytc reductase, Complex-IV reduced Cytc-Cyt oxidase, ATPIF ATP synthase inhibitory factor, UQ ubiquinone

References 1. Siegel RL, Miller KD, Jemal A (2018) Cancer statistics. CA Cancer J Clin 68(1):7–30 2. Hagmann W, Faissner R, Schnölzer M, Löhr M, Jesnowski R (2010) Membrane drug transporters and chemoresistance in human pancreatic carcinoma. Cancers 3(1):106–125 3. Ghaneh P, Costello E, Neoptolemos JP (2007) Biology and management of pancreatic cancer. Gut 56(8):1134–1152 4. Bonnet S, Archer SL, Allalunis-Turner J, Haromy A, Beaulieu C, Thompson R, Lee CT, Lopaschuk GD, Puttagunta L, Bonnet S et al (2007) A mitochondria-K+ channel axis is suppressed in cancer and its normalization promotes apoptosis and inhibits cancer growth. Cancer Cell 11(1):37–51 5. Hsu CC, Tseng LM, Lee HC (2016) Role of mitochondrial dysfunction in cancer progression. Exp Biol Med (Maywood) 241(12):1281–1295 6. Hanahan D, Weinberg RA (2011) Hallmarks of cancer: the next generation. Cell 144 (5):646–674 7. Weinberg SE, Chandel NS (2015) Targeting mitochondria metabolism for cancer therapy. Nat Chem Biol 11(1):9–15 8. Wallace DC, Fan W, Procaccio V (2010) Mitochondrial energetics and therapeutics. Annu Rev Pathol 5:297–348 9. Galluzzi L, Kepp O, Kroemer G (2012) Mitochondria: master regulators of danger signalling. Nat Rev Mol Cell Biol 13(12):780–788 10. Wallace DC (2012) Mitochondria and cancer. Nat Rev Cancer 12(10):685–698 11. Schapira AH (2012) Mitochondrial diseases. Lancet 379(9828):1825–1834 12. Verschoor ML, Ungard R, Harbottle A, Jakupciak JP, Parr RL, Singh G (2013) Mitochondria and cancer: past, present, and future. Biomed Res Int 2013:612369

6

Targeting Mitochondrial Enzymes in Pancreatic Cancer

107

13. Warburg O (1956) On the origin of cancer cells. Science 123(3191):309–314 14. Boland ML, Chourasia AH, Macleod KF (2013) Mitochondrial dysfunction in cancer. Front Oncol 3:292 15. Henze K, Martin W (2003) Evolutionary biology: essence of mitochondria. Nature 426 (6963):127–128 16. Chaffey N (2003) Alberts, B., Johnson, A., Lewis, J., Raff, M., Roberts, K. and Walter, P. Molecular biology of the cell. 4th edn. Ann Bot 91(3):401–401 17. Anderson S, Bankier AT, Barrell BG, de Bruijn MH, Coulson AR, Drouin J, Eperon IC, Nierlich DP, Roe BA, Sanger F et al (1981) Sequence and organization of the human mitochondrial genome. Nature 290(5806):457–465 18. Bayrhuber M, Meins T, Habeck M, Becker S, Giller K, Villinger S, Vonrhein C, Griesinger C, Zweckstetter M, Zeth K (2008) Structure of the human voltage-dependent anion channel. Proc Natl Acad Sci 105(40):15370 19. Anderson NM, Mucka P, Kern JG, Feng H (2018) The emerging role and targetability of the TCA cycle in cancer metabolism. Protein Cell 9(2):216–237 20. Eng C, Kiuru M, Fernandez MJ, Aaltonen LA (2003) A role for mitochondrial enzymes in inherited neoplasia and beyond. Nat Rev Cancer 3(3):193–202 21. Mitchell P, Moyle J (1967) Chemiosmotic hypothesis of oxidative phosphorylation. Nature 213 (5072):137–139 22. Mitchell P, Moyle J (1967) Acid-base titration across the membrane system of rat-liver mitochondria. Catalysis by uncouplers. Biochem J 104(2):588–600 23. Finkel T, Holbrook NJ (2000) Oxidants, oxidative stress and the biology of ageing. Nature 408 (6809):239–247 24. Drose S, Brandt U (2012) Molecular mechanisms of superoxide production by the mitochondrial respiratory chain. Adv Exp Med Biol 748:145–169 25. Salway JG (2016) Metabolism at a Glance. Wiley, Hoboken 26. Lehninger AL, Nelson DL, Cox MM, Cox MM (2005) Lehninger principles of biochemistry. Macmillan 27. Schlichtholz B, Turyn J, Goyke E, Biernacki M, Jaskiewicz K, Sledzinski Z, Swierczynski J (2005) Enhanced citrate synthase activity in human pancreatic cancer. Pancreas 30(2):99–104 28. Blum R, Kloog Y (2014) Metabolism addiction in pancreatic cancer. Cell Death Dis 5:e1065 29. Lin CC, Cheng TL, Tsai WH, Tsai HJ, Hu KH, Chang HC, Yeh CW, Chen YC, Liao CC, Chang WT (2012) Loss of the respiratory enzyme citrate synthase directly links the Warburg effect to tumor malignancy. Sci Rep 2:785 30. Wilkinson N, Pantopoulos K (2014) The IRP/IRE system in vivo: insights from mouse models. Front Pharmacol 5:176–176 31. Torti SV, Torti FM (2013) Iron and cancer: more ore to be mined. Nat Rev Cancer 13 (5):342–355 32. Yuan C, Clish CB, Wu C, Mayers JR, Kraft P, Townsend MK, Zhang M, Tworoger SS, Bao Y, Qian ZR et al (2016) Circulating metabolites and survival among patients with pancreatic cancer. J Natl Cancer Inst 108(6):djv409 33. Margittai E, Banhegyi G (2008) Isocitrate dehydrogenase: a NADPH-generating enzyme in the lumen of the endoplasmic reticulum. Arch Biochem Biophys 471(2):184–190 34. Ramachandran N, Colman RF (1980) Chemical characterization of distinct subunits of pig heart DPN-specific isocitrate dehydrogenase. J Biol Chem 255(18):8859–8864 35. Stoddard BL, Dean A, Koshland DE Jr (1993) Structure of isocitrate dehydrogenase with isocitrate, nicotinamide adenine dinucleotide phosphate, and calcium at 2.5-A resolution: a pseudo-Michaelis ternary complex. Biochemistry 32(36):9310–9316 36. Koh HJ, Lee SM, Son BG, Lee SH, Ryoo ZY, Chang KT, Park JW, Park DC, Song BJ, Veech RL et al (2004) Cytosolic NADP+-dependent isocitrate dehydrogenase plays a key role in lipid metabolism. J Biol Chem 279(38):39968–39974

108

G. Srivani et al.

37. Filipp FV, Scott DA, Ronai ZA, Osterman AL, Smith JW (2012) Reverse TCA cycle flux through isocitrate dehydrogenases 1 and 2 is required for lipogenesis in hypoxic melanoma cells. Pigment Cell Melanoma Res 25(3):375–383 38. Xu X, Zhao J, Xu Z, Peng B, Huang Q, Arnold E, Ding J (2004) Structures of human cytosolic NADP-dependent isocitrate dehydrogenase reveal a novel self-regulatory mechanism of activity. J Biol Chem 279(32):33946–33957 39. Barnes LD, Kuehn GD, Atkinson DE (1971) Yeast diphosphopyridine nucleotide specific isocitrate dehydrogenase. Purification and some properties. Biochemistry 10(21):3939–3944 40. Grzeschik KH (1976) Assignment of a gene for human mitochondrial isocitrate dehydrogenase (ICD-M, EC 1.1.1.41) to chromosome 15. Hum Genet 34(1):23–28 41. Weiss C, Zeng Y, Huang J, Sobocka MB, Rushbrook JI (2000) Bovine NAD+-dependent isocitrate dehydrogenase: alternative splicing and tissue-dependent expression of subunit 1. Biochemistry 39(7):1807–1816 42. Brody JR, Yabar CS, Zarei M, Bender J, Matrisian LM, Rahib L, Heartwell C, Mason K, Yeo CJ, Peiper SC et al (2018) Identification of a novel metabolic-related mutation (IDH1) in metastatic pancreatic cancer. Cancer Biol Ther 19(4):249–253 43. Sjöblom T, Jones S, Wood LD, Parsons DW, Lin J, Barber TD, Mandelker D, Leary RJ, Ptak J, Silliman N et al (2006) The consensus coding sequences of human breast and colorectal cancers. Science 314(5797):268–274 44. Yan H, Parsons DW, Jin G, McLendon R, Rasheed BA, Yuan W, Kos I, Batinic-Haberle I, Jones S, Riggins GJ (2009) IDH1 and IDH2 mutations in gliomas. N Engl J Med 360 (8):765–773 45. Mardis ER, Ding L, Dooling DJ, Larson DE, McLellan MD, Chen K, Koboldt DC, Fulton RS, Delehaunty KD, McGrath SD (2009) Recurring mutations found by sequencing an acute myeloid leukemia genome. N Engl J Med 361(11):1058–1066 46. Liu X, Kato Y, Kaneko MK, Sugawara M, Ogasawara S, Tsujimoto Y, Naganuma Y, Yamakawa M, Tsuchiya T, Takagi M (2013) Isocitrate dehydrogenase 2 mutation is a frequent event in osteosarcoma detected by a multi-specific monoclonal antibody MsMab-1. Cancer Med 2(6):803–814 47. Kang MR, Kim MS, Oh JE, Kim YR, Song SY, Seo SI, Lee JY, Yoo NJ, Lee SH (2009) Mutational analysis of IDH1 codon 132 in glioblastomas and other common cancers. Int J Cancer 125(2):353–355 48. Parsons D, Jones S, Zhang X, Lin J, Leary R, Angenendt P, Mankoo P, Carter H, Siu I, Gallia G, Hartigan J, Smith DR, Strausberg RL, Marie SK, Shinjo SM, Yan H, Riggins GJ, Bigner DD, Karchin R, Papadopoulos N, Parmigiani G, Vogelstein B, Velculescu VE, Kinzler KW (2008) An integrated genomic analysis of human glioblastoma multiforme. Science 321 (5897):1807–1812 49. Borger DR, Goyal L, Yau T, Poon RT, Ancukiewicz M, Deshpande V, Christiani DC, Liebman HM, Yang H, Kim H (2014) Circulating oncometabolite 2-hydroxyglutarate is a potential surrogate biomarker in patients with isocitrate dehydrogenase-mutant intrahepatic cholangiocarcinoma. Clin Cancer Res 20(7):1884–1890 50. Lu C, Ward PS, Kapoor GS, Rohle D, Turcan S, Abdel-Wahab O, Edwards CR, Khanin R, Figueroa ME, Melnick A (2012) IDH mutation impairs histone demethylation and results in a block to cell differentiation. Nature 483(7390):474 51. Dang L, White DW, Gross S, Bennett BD, Bittinger MA, Driggers EM, Fantin VR, Jang HG, Jin S, Keenan MC (2009) Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature 462(7274):739 52. Koivunen P, Lee S, Duncan CG, Lopez G, Lu G, Ramkissoon S, Losman JA, Joensuu P, Bergmann U, Gross S (2012) Transformation by the (R)-enantiomer of 2-hydroxyglutarate linked to EGLN activation. Nature 483(7390):484 53. Turcan S, Fabius AW, Borodovsky A, Pedraza A, Brennan C, Huse J, Viale A, Riggins GJ, Chan TA (2013) Efficient induction of differentiation and growth inhibition in IDH1 mutant glioma cells by the DNMT Inhibitor Decitabine. Oncotarget 4(10):1729–1736

6

Targeting Mitochondrial Enzymes in Pancreatic Cancer

109

54. Ward PS, Patel J, Wise DR, Abdel-Wahab O, Bennett BD, Coller HA, Cross JR, Fantin VR, Hedvat CV, Perl AE et al (2010) The common feature of leukemia-associated IDH1 and IDH2 mutations is a neomorphic enzyme activity converting alpha-ketoglutarate to 2-hydroxyglutarate. Cancer Cell 17(3):225–234 55. Bardella C, Pollard PJ, Tomlinson I (2011) SDH mutations in cancer. Biochim Biophys Acta 1807(11):1432–1443 56. Scheffler IE (1998) Molecular genetics of succinate:quinone oxidoreductase in eukaryotes. Prog Nucleic Acid Res Mol Biol 60:267–315 57. Zhao T, Mu X, You Q (2017) Succinate: an initiator in tumorigenesis and progression. Oncotarget 8(32):53819–53828 58. Baysal BE, Ferrell RE, Willett-Brozick JE, Lawrence EC, Myssiorek D, Bosch A, van der Mey A, Taschner PE, Rubinstein WS, Myers EN et al (2000) Mutations in SDHD, a mitochondrial complex II gene, in hereditary paraganglioma. Science 287(5454):848–851 59. Niemeijer ND, Papathomas TG, Korpershoek E, de Krijger RR, Oudijk L, Morreau H, Bayley JP, Hes FJ, Jansen JC, Dinjens WN et al (2015) Succinate dehydrogenase (SDH)-deficient pancreatic neuroendocrine tumor expands the SDH-related tumor spectrum. J Clin Endocrinol Metab 100(10):E1386–E1393 60. Selak MA, Armour SM, MacKenzie ED, Boulahbel H, Watson DG, Mansfield KD, Pan Y, Simon MC, Thompson CB, Gottlieb E (2005) Succinate links TCA cycle dysfunction to oncogenesis by inhibiting HIF-alpha prolyl hydroxylase. Cancer Cell 7(1):77–85 61. Wallace DC, Fan W (2010) Energetics, epigenetics, mitochondrial genetics. Mitochondrion 10 (1):12–31 62. Weinberg F, Chandel NS (2009) Mitochondrial metabolism and cancer. Ann N Y Acad Sci 1177:66–73 63. Chandel NS, McClintock DS, Feliciano CE, Wood TM, Melendez JA, Rodriguez AM, Schumacker PT (2000) Reactive oxygen species generated at mitochondrial complex III stabilize hypoxia-inducible factor-1alpha during hypoxia: a mechanism of O2 sensing. J Biol Chem 275(33):25130–25138 64. Selak MA, Duran RV, Gottlieb E (2006) Redox stress is not essential for the pseudo-hypoxic phenotype of succinate dehydrogenase deficient cells. Biochim Biophys Acta 1757 (5–6):567–572 65. Katagiri T, Kobayashi M, Yoshimura M, Morinibu A, Itasaka S, Hiraoka M, Harada H (2018) HIF-1 maintains a functional relationship between pancreatic cancer cells and stromal fibroblasts by upregulating expression and secretion of Sonic hedgehog. Oncotarget 9 (12):10525–10535 66. Cervera AM, Bayley JP, Devilee P, McCreath KJ (2009) Inhibition of succinate dehydrogenase dysregulates histone modification in mammalian cells. Mol Cancer 8:89 67. Launonen V, Vierimaa O, Kiuru M, Isola J, Roth S, Pukkala E, Sistonen P, Herva R, Aaltonen LA (2001) Inherited susceptibility to uterine leiomyomas and renal cell cancer. Proc Natl Acad Sci U S A 98(6):3387–3392 68. Castro-Vega LJ, Buffet A, De Cubas AA, Cascon A, Menara M, Khalifa E, Amar L, Azriel S, Bourdeau I, Chabre O et al (2014) Germline mutations in FH confer predisposition to malignant pheochromocytomas and paragangliomas. Hum Mol Genet 23(9):2440–2446 69. Sudarshan S, Shanmugasundaram K, Naylor SL, Lin S, Livi CB, O'Neill CF, Parekh DJ, Yeh IT, Sun LZ, Block K (2011) Reduced expression of fumarate hydratase in clear cell renal cancer mediates HIF-2alpha accumulation and promotes migration and invasion. PLoS One 6(6): e21037 70. Ternette N, Yang M, Laroyia M, Kitagawa M, O'Flaherty L, Wolhulter K, Igarashi K, Saito K, Kato K, Fischer R et al (2013) Inhibition of mitochondrial aconitase by succination in fumarate hydratase deficiency. Cell Rep 3(3):689–700 71. Ooi A, Wong JC, Petillo D, Roossien D, Perrier-Trudova V, Whitten D, Min BW, Tan MH, Zhang Z, Yang XJ et al (2011) An antioxidant response phenotype shared between hereditary and sporadic type 2 papillary renal cell carcinoma. Cancer Cell 20(4):511–523

110

G. Srivani et al.

72. Zheng L, Cardaci S, Jerby L, MacKenzie ED, Sciacovelli M, Johnson TI, Gaude E, King A, Leach JD, Edrada-Ebel R et al (2015) Fumarate induces redox-dependent senescence by modifying glutathione metabolism. Nat Commun 6:6001 73. Adam J, Ramracheya R, Chibalina MV, Ternette N, Hamilton A, Tarasov AI, Zhang Q, Rebelato E, Rorsman NJG, Martin-Del-Rio R et al (2017) Fumarate hydratase deletion in pancreatic beta cells leads to progressive diabetes. Cell Rep 20(13):3135–3148 74. Chen S, Chen X, Shan T, Ma J, Lin W, Li W, Kang Y (2018) MiR-21-mediated metabolic alteration of cancer-associated fibroblasts and its effect on pancreatic cancer cell behavior. Int J Biol Sci 14(1):100–110 75. Laurenti G, Tennant DA (2016) Isocitrate dehydrogenase (IDH), succinate dehydrogenase (SDH), fumarate hydratase (FH): three players for one phenotype in cancer? Biochem Soc Trans 44(4):1111–1116 76. Minarik P, Tomaskova N, Kollarova M, Antalik M (2002) Malate dehydrogenases–structure and function. Gen Physiol Biophys 21(3):257–265 77. Moreadith RW, Lehninger AL (1984) The pathways of glutamate and glutamine oxidation by tumor cell mitochondria. Role of mitochondrial NAD(P)+ dependent malic enzyme. J Biol Chem 259(10):6215–6221 78. Liu M, Chen Y, Huang B, Mao S, Cai K, Wang L, Yao X (2018) Tumor-suppressing effects of microRNA-612 in bladder cancer cells by targeting malic enzyme 1 expression. Int J Oncol 52 (6):1923–1933 79. Wang Y-P, Zhou W, Wang J, Huang X, Zuo Y, Wang T-S, Gao X, Xu Y-Y, Zou S-W, Liu Y-B et al (2016) Arginine Methylation of MDH1 by CARM1 inhibits glutamine metabolism and suppresses pancreatic cancer. Mol Cell 64(4):673–687 80. Hirst J, King MS, Pryde KR (2008) The production of reactive oxygen species by complex I. Biochem Soc Trans 36(Pt 5):976–980 81. Chatterjee A, Mambo E, Sidransky D (2006) Mitochondrial DNA mutations in human cancer. Oncogene 25(34):4663–4674 82. Giorgio V, von Stockum S, Antoniel M, Fabbro A, Fogolari F, Forte M, Glick GD, Petronilli V, Zoratti M, Szabo I et al (2013) Dimers of mitochondrial ATP synthase form the permeability transition pore. Proc Natl Acad Sci U S A 110(15):5887–5892 83. Kim T, Thu VT, Han IY, Youm JB, Kim E, Kang SW, Kim YW, Lee JH, Joo H (2008) Does strong hypertrophic condition induce fast mitochondrial DNA mutation of rabbit heart? Mitochondrion 8(3):279–283 84. Shidara Y, Yamagata K, Kanamori T, Nakano K, Kwong JQ, Manfredi G, Oda H, Ohta S (2005) Positive contribution of pathogenic mutations in the mitochondrial genome to the promotion of cancer by prevention from apoptosis. Cancer Res 65(5):1655–1663 85. Adam J, Yang M, Soga T, Pollard P (2014) Rare insights into cancer biology. Oncogene 33 (20):2547

7

Diabetes with Pancreatic Ductal Adenocarcinoma Gowru Srivani, Begum Dariya, Afroz Alam, and Ganji Purnachandra Nagaraju

Abstract

Diabetes and pancreatic ductal adenocarcinoma (PDAC) are common diseases and affect the same organ, pancreas. PDAC has a poor prognosis and response to conservative therapy. Diabetes is recently been correlated with mortality and morbidity from PDAC. The association between diabetes and PDAC stems from the structural association between the endocrine and exocrine pancreas and aberrant expression of hormones from islets. It can also result from other etiological factors including stress, inflammation, smoking, alcohol consumption, change in the diet, as well as inherited syndromes that affect PDAC tissue. Epidemiological evidence suggests that diabetes increases the risk for PDAC development. Insulin resistance, hyperinsulimenia, hyperglycemia, chronic inflammation, and their elementary mechanisms can contribute to the development of diabetes-associated PDAC. Signal transduction pathways that regulate metabolic functions also play a crucial role in the development of PDAC, promoting tumor proliferation, cell growth, differentiation, angiogenesis, and metastasis. In another way, PDAC is also a causative factor for diabetes, although the mechanisms are not well understood. Effective biomarkers might thus help detect the increased risk of PDAC. Furthermore, greater understanding of the pathological mechanisms linking diabetes to PDAC could guide the development of new therapeutic agents to prevent diabetes associated with PDAC.

G. Srivani · B. Dariya · A. Alam Department of Bioscience and Biotechnology, Banasthali University, Vanasthali, Rajasthan, India G. P. Nagaraju (*) Department of Hematology and Medical Oncology, Winship Cancer Institute, Emory University, Atlanta, GA, USA e-mail: [email protected] # Springer Nature Singapore Pte Ltd. 2019 G. P. Nagaraju, A. BM Reddy (eds.), Exploring Pancreatic Metabolism and Malignancy, https://doi.org/10.1007/978-981-32-9393-9_7

111

112

G. Srivani et al.

Keywords

PDAC · Type 2 diabetes · Exocrine · Endocrine · Hyperglycemia · Hyperinsulinemia

Abbreviations 4-HNE Akt AMP AMPK ATP ASK-1 CI COX2 DNA ERK ETC FGD-PET FTZ-F1 GI GLUT GWAS HNF-3β IER IGFBP-1 IGF IGFR IKK IL-6 IL-8 IR IRS JNK LKB1 LOOH LRH1 MDA MEK MMP-7 mTOR NADPH NF-κB

4-hydroxyl-2-nonenal Protein kinase B Adenosine monophosphate AMP-activated protein kinase Adenosine triphosphate Apoptosis signaling kinase-1 Confidence interval Cyclooxygenase Deoxyribonucleic acid Extracellular signal-regulated kinases electron transport chain F-18-Fluoro-deoxyglucose (FDG)-positron emission tomography (PET) Fushi-tarazu factor-1 Glycemic index Glucose transporter Genome-wide association studies Hepatocyte nucleoside factor-3β Intermittent energy restriction Insulin growth factor-binding protein-1 Insulin growth factor Insulin growth factor receptors Inhibitor of kB kinase Interleukin-6 Interleukin-8 Insulin receptor Insulin receptor substrate 1 c-Jun N-terminal kinases Liver kinase B1 Lipid hydroperoxides Liver receptor homolog-1 Malondialdehyde Mitogen-activated protein kinase Matrix metalloproteinase-7 Mammalian target of rapamycin Nicotinamide adenine dinucleotide Nuclear factor kappa B

7

Diabetes with Pancreatic Ductal Adenocarcinoma

NO NR5A2 PDAC PDX-1 PI3K RIP-1 RO ROS RR SO4 SODD STAT TNF-α TRADD VAT VEGF

7.1

113

Nitric oxide Nuclear receptor superfamily member Pancreatic ductal adenocarcinoma Pancreatic duodenal homeobox Phosphatidylinositol 3 kinase Receptor interacting protein Alkoxyl radical Reactive oxygen species Relative risk Sulfate radical Silencer of death domain Signal transducer and activator of transcription 3 Tumor necrosis factor-α TNF receptor-associated death domain Vascular adipose tissue Vascular endothelial growth factor

Introduction

Pancreatic ductal adenocarcinoma (PDAC) is the most lethal cancer and fourth cause for cancer-related mortalities in the world. Its prevalence and mortality rates of PDAC are almost equal (~ 38,000 deaths and 45,000 new cases in the USA, 2013) [1]; it has a poor survival rate of 6 months, and survival for 5 year is less than 8%. Ninety-five percent of pancreatic malignancies arise in the pancreatic exocrine glands [2, 3] leading to PDAC and are usually diagnosed at the head and neck region of the pancreas [2].The life span risk of PDAC is 1% in 65-year-old patients and prevalence increases with age [4]. PDAC is predicted to become the second leading cause for cancer mortality in the USA by 2030 [5]. It is mainly caused due to the lifestyle modifications including exposure to tobacco, irregular diet habits, high alcohol consumption, chronic stipulation (chronic pancreatitis and chronic inflammation), obesity, and diabetes (Fig. 7.1) [1]. Among these factors, long-standing diabetes can enhance the risk of PDAC development. Although the connection between PDAC and diabetes is biologically complex, it has been thoroughly investigated to elucidate its mechanisms of action. There seems to be a bidirectional association between diabetes and PDAC [6]. For instance, recent epidemiological studies have proven that the risk of developing different cancers (PDAC, breast, colorectal, liver, urinary bladder, endometrial) is increased in patients with elevated levels of diabetes [7]. This is due to the fact that both cancer and diabetes have related risk factors (RR, 1.5–2.0) including obesity, diet, heredity, and being physically inactive, which elevate the incidence of PDAC in patients [8]. Moreover, diabetes associated with PDAC is detected 3 years prior to the development of PDAC and is thus considered as a risk factor (1.73–1.94) for PDAC [9].

114

7.2

G. Srivani et al.

Evidence for Diabetes as a Risk Factor for PDAC

Everhart et al. have observed a close relation between diabetes and PDAC in their 30 epidemiological studies. Using 20 studies for meta-analysis [10], they found that the pooled RR is 2.1 for diabetes within the time period of 1 year prior to PDAC diagnosis, while the death of pooled RR was 2.1 for diabetes for a time period of 5 years [10]. Finally, authors emphasized that PDAC can be considered as one of the complications of diabetes [10]. Furthermore, another study analyzing 35 cohort studies described an association between diabetes and PDAC [11], on the metaanalysis basis of 27 studies to evaluate the relative risk on prevalence and mortality. These results confirmed that high risk of PDAC was identified in less than 1 year of diagnosing diabetes (i.e., relative risk 5.4; 95% confidence interval (CI); 3.49–8.30) [11]. This data shows that a shorter duration is equal to a higher risk at 1.5-fold after 5–9 years [11]. The authors strongly concluded that diabetes is an early indicator for PDAC and that newly diagnosed diabetes patients should be extremely attentive to PDAC development [11]. In the USA, a case-control study based on the population conducted in 2153 controls with 526 prevalence cases significantly revealed that diabetes shows an optimistic trend (P ¼ 0.016) in high-risk patients with increasing years before cancer diagnosis. Risk (OR 1.3; 95% CI, 0.4–4.0) of onset of diabetes slightly increased within 1 year of cancer diagnosis [12]. A larger study on meta-analysis was conducted on pooled data from three casecontrol studies including 2192 incidence cases and 5113 controls of diabetes Diabetes

Inflammation

Insulin resistance Mutagens

Hyperinsulimenia Hyperglycemia Loss of physical activity Metabolic syndrome Fructuose

Susceptivity Family history Inheritated syndrome

Metabolites of alcohol Environmental carcinogens Processed food Ex.Red meat Saturated animal fat

Excess alcohol consumption Exposure to tobacco Obesity Pancreatitis Helicobacter pylori

PDAC

Fig. 7.1 The main causative factors for PDAC development and progression from diabetes. The main risk factors are hyperinsulimenia, metabolic syndrome, obesity, excess alcohol abuse, and helicobacter pylori

7

Diabetes with Pancreatic Ductal Adenocarcinoma

115

associated with PDAC [13]. The findings revealed that PDAC risk was reduced as the duration of diabetes increased. PDAC risk estimated as per pooled data was as follows: individuals diagnosed with diabetes less than 2 years versus individuals with diabetes diagnosed greater than 2 years prior to diagnosis had ORs with 95% CIs of 8 years showed long duration of low plasma IGFBP-1 and higher risk for PDAC (RR 3.30, 95% CI, 1.48–7.35). Hence, these results indicate that low plasma IGFBP-1 levels significantly promote PDAC risk. Another study investigated the IGF-1 gene polymorphism to illuminate the underlying processes that mediate the correlation existing between diabetes and PDAC development. Results from this study revealed an association between the IGF-1 gene polymorphic variant IGF-1 3’UTR Ex4 (G > C) C allele, diabetes, and increased risk of PDAC development [31]. Finally, evidence from epidemiological

118

G. Srivani et al.

studies indicates that the associations between long-standing diabetes, insulin resistance, and IGF-1 are important targets for conventional treatment of PDAC.

7.6

Inflammation

Evidences from various studies suggest that inflammation plays a crucial role in insulin effects, and type 2 diabetes upregulates the signaling pathways leading to PDAC development. A recent review suggests that adipose tissue in endocrine organs can regulate the release of fatty acids, adiponectin, leptin, PAL-1, VEGF, and proinflammatory cytokines including TNF-α, IL-6, and resistin [32]. Dysfunctions in the adipose tissue may thus play a vital role in the progression of diabetes and various cancers. Increased levels of proinflammatory cytokines released by dysregulated adipose tissue can promote tumorigenesis [33]. TNF-α acts as a key regulatory molecule in innate immunity and regulates apoptosis under physiological conditions. Increased levels of TNF-α bind to TNFRS, resulting in the production of inhibitory proteins such as SODD (silencer of death domain) and TRADD (TNF receptor-associated death domain) that bind to the extra adapter protein RIP-1 (receptor-interacting protein). This leads to the formation of the TRADD-RIP1-TRAF complex, which initiates various signal transduction pathways including the MAP3K cascade (ASK-1: apoptosis signaling kinase-1). This cascade is linked to the TRADD-RPI1-TRAF complex that activates MAP2Ks, MEK-4, and MEK-6, enhancing the expression of JNK (c-Jun N-terminal kinases) and TPL 2-MEK-ERK pathways and TNF-α-PAI-1/VEGF. RIP1 engages with MEKK-3 and TGF-β–TAK1 to activate the IKK (Inhibitor of kB kinase) complex, which undergoes phosphorylation leading to the ubiquitination of IkBα. This results in the release of free NF-kB subunits, subsequently translocated to the nucleus, and induces the gene transcription involved in cell proliferation and survival [34–36]. IL-6 plays a critical role in the inflammation reaction and B-cell maturation under physiological conditions. Elevated IL-6 levels resulted from the JAK–STAT3 signaling cascade activation induced cell cycle progression and increased the cell survival and invasion [36, 37]. These pathways contribute to PDAC development by diminishing immune function. Recently, Hertzer et al. [38] emphasized in their in vivo study on high-fat, high-calorie diets that induced Kras G12 mice displayed elevated levels of VAT (vascular adipose tissue), and obesity and inflammation were correlated with the highest risk for PDAC development. These results strongly propose that adipose tissue may serve as a therapeutic target to treat and prevent PDAC. Furthermore, studies investigating energy homeostasis support epidemiological data indicating a correlation between obesity, diabetes, and cancer death. For instance, few studies have shown that tumor cells can become more aggressive in overfed animals when compared to animals fed with controlled caloric diets [39, 40]. These studies show that caloric restriction can mediate intermittent energy restriction (IER) and inhibit the inflammation mediated-insulin effect of the diet on

7

Diabetes with Pancreatic Ductal Adenocarcinoma

119

tumor development. They also specify the variation between tumors with respect to the signaling cascade involved and the type of diet that influences tumor progression. Moreover, these results suggest that IGF and IGF signaling can constitute therapeutic targets for the development of novel therapeutic strategies against various diseases including PDAC [40, 41].

7.7

Oxidative Stress

The evidences also revealed that insulin, IGF-1, and type 2 diabetes may promote increased risk for PDAC by enhanced oxidative stress [42]. This is mainly due to the increased release of free radicals in the mitochondria, including ROS (reactive oxygen species), that disrupt the functions of DNA, proteins, and lipids, resulting in imbalance of metabolism [43]. High oxidative stress can release large amounts of fatty acid peroxidases that are expressed in high toxicity mutagenic functions. These compounds are (MDA) malondialdehyde and 4-HNE (4-hydroxyl-2-nonenal), both released in the adipose tissue leading to dysfunctions of the metabolic system [44]. This oxidative stress is mainly caused by imbalances between antioxidants and free radicals, especially ROS. ROS are key regulatory molecules that can disrupt gene transduction and act as intermediate molecules in signaling pathways. They can affect mitogenic functions, viz., cell growth and survival, through growth factor receptors that stimulate cell migration and can alter the tumor microenvironment by inducing angiogenesis and inflammation. In fact, tumor cells adapt and maintain their ROS levels to escape from apoptosis [45]. In PDAC, NADPH (nicotinamide adenine dinucleotide) is the major factory for intracellular ROS generation. Most of the enzymes in the mitochondria are engrossed in oxidative stress. Numerous varieties of free radicals can contribute to oxidative stress, including peroxynitrite, nitric oxide (NO), alkoxyl radical (RO), lipid hydroperoxides (LOOH), sulfate radical (SO4), and nitrogen-centered radical and metal-oxygen complexes. In addition, other cell components are also engaged in the internal oxidative stress, viz., xanthine oxidase, peroxisome, and their enzymes, especially P450 complex-detoxifying enzymes and NADPH oxidase complex, which contains NOX family [46]. Along with these free radicals, ROS generates various redox agent groups that play a vital role in many intra- and extracellular processes [47]. ROS can also enhance various signaling cascades essential for triggering tumor progression via cell growth, proliferation, angiogenesis, invasion, and apoptosis. This includes cell differentiation mediated through MAPK, via NF-kB and ERK1/2 [48]; evasion of cell death by stimulation of JAK/STAT, c-SRC, and PIK3/Akt signal cascade [49]; angiogenesis mediated by the generation of VEGF and angiopoietin [50]; and metastasis and tissue incursion by matrix metalloproteinases (MMP) inflection into the extracellular matrix [50]. Indeed, various epidemiological studies have unveiled a connection between oxidative stress and insulin resistance-mediated diabetes with PDAC. Studies additionally indicate that oxidative stress activates the initiation of signaling events that promote

120

G. Srivani et al.

inflammation and insulin resistance, leading to tumor proliferation, growth survival, and adaptation of metabolic activities to apoptosis evasion [42, 51, 52]. In PDAC, hypoxia is a distinguishing factor. In fact, the release of HIF by ROS induces PIK3/Akt pathways [53]. Furthermore, ROS can hinder autophagy through the upregulation of the signaling Akt/mTOR pathway. This data suggests the ROS can induce malignancy and is capable of promoting sensitivity to chemo drugs by inhibiting the mTOR pathway [54]. Some experimental studies have shown that antioxidant supplementation (α-lipoic acid or vitamin E) may reduce the oxidative stress caused by insulin resistance [55]. Inducing insulin resistance in the experimental models by feeding them high sucrose and fructose diet has revealed a direct relation between insulin resistance and oxidative stress. In this study, the postprandial time period showed that hyperglycemia might cause oxidative stress through different signaling mechanisms that increase ROS and superoxides by the ETC (electron transport chain) [56]. This leads to the cellular redox state, downregulates tyrosine phosphorylation, and upregulates the insulin receptor substrate 1, thus suppressing the insulin signaling pathway [56].

7.8

Genomic Associations of Diabetes and PDAC

In diabetes and PDAC, heterogeneity is a common feature. The epidemiological and clinical studies indicate that diabetes increases PDAC risk. Although the susceptibility of genetic variants is extensively different, there is a modest overlap. Interestingly, both diabetes and PDAC share common features: i) Etiological risk factors, i.e., obesity, exposure to smoke, and alcohol consumption, may influence the genetic factors to increase risk. ii) Segregation of Mendelian analysis shows that a number of families display hereditary patterns. iii) Diagnosing at different age periods shows the related risk in some families [57]. GWAS (genome-wide association studies) have revealed that certain unpredicted genetic variants might change the risk for causing diabetes, and few diabetes susceptibility gene loci are often involved in tumor differentiation and progression [58]. In diabetes, there are 50 gene variants identified as genetic risk factors with minimal effect, including PPPARG, TCF7L2, KDNJ11, HDF1B, FTO, SLC20A8, and WFS1, and are detected via GWAS analysis [59]. Interestingly, GWAS analysis has identified one of the PC susceptibility genes, namely NR5A2, also called as LRH1 (liver receptor homolog-1). NR5A2 is one of the orphan nuclear receptor superfamily members that encodes the nuclear receptor subfamily gene FTZ-F1 (fushi-tarazu factor-1). It is highly expressed in the liver, intestine, exocrine pancreas, and ovaries [60, 61]. Functionally, it plays a vital role in developing digestive organs, steroidogenesis, cholesterol, bile acid synthesis, inflammation regulation in gut and liver, and cell proliferation in certain diseases including PDAC [62–64]. Studies have revealed that NR5A2 overexpression can promote cell growth, proliferation, and metastasis in PDAC [65, 66]. Thus, it might be considered as a potential target for PDAC. NR5A2 regulates the expression of several genes, which

7

Diabetes with Pancreatic Ductal Adenocarcinoma

121

play a crucial role in the development of pancreas, insulin secretion, and β-cell function. These include hepatocyte nucleoside factor-3β (HNF-3β), HNF-1α, HNF-4α, and PDX-1. Conversely, NR5A2 expression is controlled by HNF-3β and HNF-1 [67]. Among these, PDX-1 (pancreatic duodenal homeobox) is an important regulator of endocrine pancreas development and is also involved in maintaining the endocrine pancreas functions through triggering gene transcription of islet amyloid polypeptide, insulin, glucokinase, GLUT-2, and somatostatin [67, 68]. Mutations in the HNF-3β and PDX-1 genes can cause type 2 diabetes and maturity-onset diabetes [69]. PDX-1 overproduction in pancreas upregulates the expression of MMP-7 in PDAC leading to tumor cell proliferation and invasiveness [70]. Moreover, the mutation in K-ras gene in PDAC may promote the overexpression of PDX-1, resulting in tumor cell proliferation, invasiveness, and survival [71]. Furthermore, cells undergoing embryogenesis and organogenesis require PDX-1 for survival and development of endocrine pancreas, but the overexpression of PDX-1 leads to dysregulation of islet cells through enhanced PI3K/AKT- and RAS-mediated signaling, leading to PDAC development, including proliferation, invasiveness, angiogenesis, and survival [71]. A better understanding of the NR5A2 gene significance and underlying mechanisms of HNF and PDX-1 in the development of PDAC may serve as potential molecular biomarker to diagnose and treat diabetes-associated PDAC.

7.9

Antidiabetic Therapy as a Risk Factor for PDAC

Controlling glucose levels is the major objective of effective diabetic maintenance, which leads to diminish mortality and morbidity by decreasing the risk of diabetesassociated PDAC. Clinicians consider numerous factors when choosing diabetic medications such as the type of diabetes, chronic adverse effects (including hyperglycemia, weight gain, fluid retention, gastrointestinal intolerance), patient characteristics, comorbidities, and cost of the treatment. Type 1 diabetic patients account for ~5–10% of the diabetic population globally. The deficiency of insulin affected due to autoimmune damage of the pancreatic β-cells requires life-long insulin therapy. In contrast, type 2 diabetes is frequent and is the reason for 95% of the diabetic population worldwide. However, 80% of type 2 diabetes is connected with decreased physical activity and obesity and usually develops from the prediabetic stage, distinguished by hyperinsulinemia (insulin resistance). Furthermore, it is always accompanied by reduced insulin release. Hence, loss of insulin increases both during fasting and post meal hyperglycemia [72]. Continued loss of insulin production and decreased incretin effects as well as various pathophysiological deficiencies lead to hyperglycemia of type 2 diabetes, which requires increased insulin therapy. The selection of the pharmacological agent prescribed to the patient is based on clinical symptoms and ongoing risk factors. Singh et al. [73] conducted a meta-analysis on 11 studies (6 cohort, 3 casecontrol, and 2 randomized control trails (RCTs)) and reported 1770 PDAC cases in

122

G. Srivani et al.

730,664 patients suffering with diabetes on several antidiabetic medications. This analysis shows a significant relationship between insulin, metformin, and sulfonylureas use and the risk of developing PDAC in diabetic patients. Metformin users did not show PDAC risk in diabetes patients (P ¼ 0.073) as compared with non-users of metformin. In contrast, the use of sulfonylureas showed a significant associated risk of PDAC in diabetic patients [73].

7.10

Metformin

Numerous diabetic studies have investigated treatment with biguanide metformin, an antidiabetic drug frequently prescribed alone or in combination therapy. The risk of cancer-related mortality is decreased by metformin therapy when compared to insulin and insulin secretagogues treatment [74]. An epidemiologic meta-analysis of two studies was carried out to evaluate the outcome of metformin on cancer prevalence and death rate in patients with diabetes. The results showed a 32% decrease with overall summary of RR for cancer as 0.69 and 95% for CI (confidence interval) as 0.61–0.79 following metformin treatment as compared to other antidiabetic drugs [75]. Furthermore, a meta-analysis of a retrospective cohort of 62,809 patients with diabetes [76] and another case-control study on 973 patients with PDAC versus 863 case controls [77] were conducted separately with same time period. These studies reveled that metformin users display reduced risk of PDAC compared to sulfonylurea and insulin users. The glucose levels are reduced by metformin by decreasing the production of hepatic glucose and enhancing the insulin sensitivity. It slightly elevates plasma insulin concentration and suppresses the growth-regulating effect of insulin and IGFs [78]. In diabetic patients, metformin treatment significantly induces endothelial function, such as reduced plasma levels of soluble vascular cell adhesion molecule 1, vascular endothelial growth factor, von Willebrand factor, and soluble E-selectin tissue-type plasminogen inhibitor and activator [78]. Metformin suppresses the proinflammatory cytokines (IL-6 and IL-8) through the inhibition of NF-kB activation by the PI3k-Akt pathway [79]. These molecules play a crucial role in fibrosis, thrombosis, and angiogenesis which are essential for the development and progression of PDAC [78].

7.11

Molecular Mechanism of Metformin

Metformin acts mainly by inhibiting hepatic gluconeogenesis, thus reducing circulating glucose levels, increasing insulin sensitivity, and reducing the intestinal absorption of glucose [80]. Various mechanisms have been suggested to elucidate the inhibitory function of metformin on hepatic gluconeogenesis and glucose uptake such as suppression of the mitochondrial respiratory chain complex and AMPK activation (AMP-activated protein kinase) signaling pathway [81].

7

Diabetes with Pancreatic Ductal Adenocarcinoma

123

AMPK is a master controller of several metabolic pathways and can inhibit glucose production in hepatocytes through liver kinase B1 (LKB1). AMPK activation by LKB1 is necessary for mediating the blood glucose effect of metformin [82]. A study performed by Shaw et al. (2005) observed that loss of LKB1 leads to phosphorylation and increased activation of AMPK, suggesting that treatment with metformin can enhance the activity of AMPK and decrease blood glucose levels in a LKB1-dependent process [83]. ATP production is mainly carried out in the mitochondria by oxidative phosphorylation, such as catabolic mechanism, namely, amino acids and fatty acids oxidation. Metformin treatment inhibits mitochondrial electron transport chain complex-I to enhance the activation of AMPK [81, 84]. This decreases cellular ATP production and elevates AMP levels, leading to increased glucose uptake and glycolysis [84]. AMPK contributes to cellular energy (ATP) levels and is activated by low intracellular ATP concentrations. Metformin suppresses the mTORC1 pathway through AMPK [85], which responds to energy stress by inhibiting cell proliferation and protein synthesis, suppressing the mTOR1 pathway [86]. AMPK plays a crucial role as a metabolic checkpoint for harmonizing energy levels and cell growth to ensure the initiation and maintenance of normal cell division and polarity [87]. In another study, He et al. [88] revealed that metformin suppresses hepatic gluconeogenesis by phosphorylating CBP (CREB-binding protein) at serine 436 through AMPK-PKCi/l, resulting in the separation of the CREB-CBP-CRTC2 transcription complex and downregulating gluconeogenic genes. Pearce et al. [89] showed in their study that long-term therapy of metformin increases CD8 T-cell half-life and improves the efficacy of anticancer vaccination. Metformin promotes apoptosis in cancer cells through the LKB1/AMPK pathway. This pathway regulates the phosphorylation of p27 at Thr 198 to stabilize p27, which permits cells to survive growth factor withdrawal and metabolic stress by autophagy [90] (Fig. 7.2). The antitumor activity of metformin not only regulates hormonal and metabolic functions but also regulates cell division and immune function.

7.12

Metformin and Its PDAC Preventive Effects

Epidemiologically, antidiabetic medications have been associated with decreased incidence of PDAC recurrence and death in diabetes-related PDAC patients. Recently, Li et al. [91] performed meta-analysis studies on nine retrospective cohort studies and two randomized controlled trials. These studies showed a considerable development in survival (HR, 0.86%; 95% CI, 0.76–0.97, P < 0.05) in metformin users as compared with controls. This study suggests that the metformin effect depends on the stage of tumor, with noticeable improvement in the survival rate of patients with locally advanced diseases, however, not with metastatic PDAC patients [91]. Another study by Kisfalvi et al. [92] suggested that metformin significantly suppressed PDAC growth via disturbing the signaling crosstalk between GPCR (G protein-coupled receptor) and IGF (insulin-like growth factor) receptor. Furthermore, oral metformin administration showed that metformin can significantly reduce

124

G. Srivani et al.

PANC-1 and MIA PaCa-2 cell growth, xenografted on the flank nude cells. Hence, these results indicate that metformin could be a potential preventive agent for human PDAC. Additionally, Li et al. [77] revealed that the risk of progressing PDAC in metformin users decreased by 62% (OR, 0.38; CI, 95%, 0.22–0.69; P ¼ 0.001), as compared with the non-users of metformin. Another retrospective study performed by Sadeghi et al. [93] revealed that PDAC patients with diabetes show improved survival when treated with metformin [93]. Metformin has also been shown to check the endorsement effect of high fat diet on N-nitrosobis (2-oxopropyl) amine-induced pancreatic carcinogenesis in Syrian hamsters [94] and to suppress PANC-1 and MIA PaCa-2 tumor growth in xenograft models using athymic nude mice [95]. Furthermore, metformin inhibited PDAC growth by various inflammatory and tumorigenesis signaling pathways. For instance, Tan et al. [96] showed in their study that metformin treatment significantly reduced tumor growth, volume, burden, and weight by downregulating the NF-κB-STAT3 signaling pathways and activating AMPK. In another study, Metformin suppressed angiogenesis by inhibiting VEGF release in the tumor microenvironment and reducing tumor neovascularization. Additionally, it increased the chemosensitivity of gemcitabine by inactivating pancreatic stellate cells in PDAC [97]. These findings illustrate the antitumor activity of metformin in preventing diabetes-associated PDAC and suggest a novel approach for PDAC therapy and prevention. Insulin secretagogues, such as sulfonylureas, are rapid acting glinides [8]. Sulfonylureas are crucial for raising the plasma insulin levels but effectively act when pancreatic β-cells are present [98]. Moreover, they induce β-cells to release plasma insulin leading to β-cell depolarization [8]. For the past 50 years, sulfonylureas, for example, glipizide, glyburide, and glimepiride, have been used to treat type 2 diabetes patients. While sulfonylurea drugs can effectively reduce blood glucose concentrations, they cause weight gain and hyperglycemia [8].

7.13

Molecular Mechanism of Sulfonylureas

Sulfonylureas increase the plasma insulin concentrations by blocking the ATP-sensitive K-channels in the pancreatic plasma membrane. The release of insulin is initiated by a series of events. In pancreatic β-cells, the (potassium) K+-channel plays a vital role in regulating insulin secretion in order to respond to sulfonylureas and nutrient secretagogues. The K+-channel is mainly located in smooth muscles, skeletal muscle, and β-cells of the pancreas, kidney epithelia, and neurons. Plasma insulin levels can be increased through two different mechanisms: first, physiological stimulation through pancreatic β-cells and, second, reduction in hepatic clearance of insulin. Given the secretary function of sulfonylureas, it acts by binding to specific receptors on pancreatic β-cells. Stimulated insulin secretion raises blood glucose levels. Glucose enters into the pancreatic β-cells where it is metabolized by glycoly-

7

Diabetes with Pancreatic Ductal Adenocarcinoma

125

sis and mitochondria, producing ATP and leading to increased intracellular ATP levels. This provokes the closure of ATP-dependent K+-channels in the plasma membrane of β-cells, thus decreasing the membrane’s K-permeability and causing the depolarization of the cellular membrane. This in turn leads to the opening of voltage-dependent Ca 2+ channels that enhance Ca 2+ inflow to the cytosol, thus triggering the exocytosis of insulin. This process is mainly caused by the actomyosin contraction and mediates the release of large amounts of insulin (Fig. 7.2) [98, 99]. A number of studies have shown increased risk of cancer and cancer-related mortality among patients with diabetes treated with sulfonylurea as compared with patients treated with metformin and other drugs [77, 100, 101]. It is uncertain whether this greater risk is associated with the effects of sulfonylurea or the more protective effect of metformin or if it is due to the effects associated with therapy selection and cancer risk [100]. Therefore, studies evaluating the duration, age, dose, and persistence of use are required to examine the association with particular cancer sites. The tumor-inducing effects of increased insulin levels and resulting IGF-1, insulin use in diabetic patients, and its association with greater risk of cancer have harnessed a wide attention [102, 103]. The diabetic patients using insulin for longer periods have been associated with increased risk of various cancers, including PDAC; however, the patients who depend for short period of insulin dependency say for less than 3 years are also at higher risk [104, 105]. In a study evaluating pooled data from three large case-control studies, long-term (> 10 years) insulin users had reduced risk as compared with short-time insulin users. Therefore, shorttime (3–10 years of insulin) users showed a 20% increased risk compared to patients that never used insulin [13]. In another case study of PDAC patients, more than 80% of patients using insulin only started taking it at least 2 years prior to their cancer diagnosis, which resulted from diabetes causing PDAC or long-standing diabetes provoked by the PDAC [77]. These above studies shown that PDAC risk in diabetic patients is correlated with a reduced number of patients using insulin. Therefore, a larger study sample is required to review the feasible effect of insulin use on PDAC risk.

7.14

Conclusion

The association between diabetes and PDAC is biologically complex. The risk of PDAC progression is higher in diabetic patients, with significant associated factors that modulate metabolic functions. Greater understanding of these mechanisms may lead to prognostic biomarkers for the early detection of PDAC. The protective effect of the antidiabetic drugs against PDAC provides an opportunity for intervention and prevention of this cancer (Fig. 7.3).

126 Fig. 7.3 Molecular mechanism of sulfonylureas: insulin secretion from pancreatic β-cells. Sulfonylureas increases the plasma insulin levels by inhibiting the ATP-dependent K+-channels through Ca 2+ channels-mediated exocytosis process

G. Srivani et al.

Sulfonylureas Binds to high affinity of sulfhonylaurea receptor on pancreatic β – cell Closures the influx of ATP dependent K+ - channels β – cell depolarization opens the voltage –dependent Ca 2+ channels

Increased Ca 2+ inflow in to the cytosol exocytosis Secreted large amount of insulin

References 1. Becker AE, Hernandez YG, Frucht H, Lucas AL (2014) Pancreatic ductal adenocarcinoma: risk factors, screening, and early detection. World J Gastroenterol 20(32):11182–11198 2. Saad AM, Turk T, Al-Husseini MJ, Abdel-Rahman O (2018) Trends in pancreatic adenocarcinoma incidence and mortality in the United States in the last four decades; a SEER-based study. BMC Cancer 18(1):688–688 3. Siegel RL, Miller KD, Jemal A (2015) Cancer statistics, 2015. CA Cancer J Clin 65(1):5–29 4. Howlader N, Noone A, Krapcho M, Garshell J, Neyman N, Altekruse S (2013) SEER cancer statistics review, 1975–2010. National Cancer Institute, Bethesda 5. Rahib L, Smith BD, Aizenberg R, Rosenzweig AB, Fleshman JM, Matrisian LM (2014) Projecting Cancer Incidence and Deaths to 2030: The Unexpected Burden of Thyroid, Liver, and Pancreas Cancers in the United States. Cancer Res 74:2913 6. Sah RP, Nagpal SJS, Mukhopadhyay D, Chari ST (2013) New insights into pancreatic cancerinduced paraneoplastic diabetes. Nat Rev Gastroenterol Hepatol 10(7):423–433 7. Vigneri P, Frasca F, Sciacca L, Pandini G, Vigneri R (2009) Diabetes and cancer. Endocr Relat Cancer 16(4):1103–1123 8. Giovannucci E, Harlan DM, Archer MC, Bergenstal RM, Gapstur SM, Habel LA, Pollak M, Regensteiner JG, Yee D (2010) Diabetes and cancer: a consensus report. Diabetes Care 33 (7):1674–1685 9. Aggarwal G, Kamada P, Chari ST (2013) Prevalence of diabetes mellitus in pancreatic cancer compared to common cancers. Pancreas 42(2):198–201 10. Everhart J, Wright D (1995) Diabetes mellitus as a risk factor for pancreatic cancer: a metaanalysis. JAMA 273(20):1605–1609 11. Ben Q, Xu M, Ning X, Liu J, Hong S, Huang W, Zhang H, Li Z (2011) Diabetes mellitus and risk of pancreatic cancer: a meta-analysis of cohort studies. Eur J Cancer 47(13):1928–1937

7

Diabetes with Pancreatic Ductal Adenocarcinoma

127

12. Silverman DT (2001) Risk factors for pancreatic cancer: A case-control study based on direct interviews. Teratog Carcinog Mutagen 21(1):7–25 13. Li D, Tang H, Hassan MM, Holly EA, Bracci PM, Silverman DT (2011) Diabetes and risk of pancreatic cancer: a pooled analysis of three large case-control studies. Cancer Causes Control 22(2):189–197 14. Himsworth H, Kerr R (1939) Insulin-sensitive and insulin-insensitive types of diabetes mellitus. Clin Sci 4:119–152 15. Vander Heiden MG, Cantley LC, Thompson CB (2009) Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science (New York, NY) 324 (5930):1029–1033 16. Yun J, Rago C, Cheong I, Pagliarini R, Angenendt P, Rajagopalan H, Schmidt K, Willson JKV, Markowitz S, Zhou S et al (2009) Glucose deprivation contributes to the development of KRAS pathway mutations in tumor cells. Science (New York, NY) 325(5947):1555–1559 17. Hine RJ, Srivastava S, Milner JA, Ross SA (2003) Nutritional links to plausible mechanisms underlying pancreatic cancer: a conference report. Pancreas 27(4):356–366 18. Mulholland HG, Murray LJ, Cardwell CR, Cantwell MM (2008) Glycemic index, glycemic load, and risk of digestive tract neoplasms: a systematic review and meta-analysis. Am J Clin Nutr 89(2):568–576 19. Reaven GM (1993) Role of insulin resistance in human disease (syndrome X): an expanded definition. Annu Rev Med 44(1):121–131 20. Lakka H-M, Laaksonen DE, Lakka TA, Niskanen LK, Kumpusalo E, Tuomilehto J, Salonen JT (2002) The metabolic syndrome and total and cardiovascular disease mortality in middleaged men. JAMA 288(21):2709–2716 21. Giovannucci E, Harlan DM, Archer MC, Bergenstal RM, Gapstur SM, Habel LA, Pollak M, Regensteiner JG, Yee D (2010) Diabetes and cancer: a consensus report. CA Cancer J Clin 60 (4):207–221 22. Pollak M (2008) Insulin and insulin-like growth factor signalling in neoplasia. Nat Rev Cancer 8(12):915 23. Singh P, Alex JM, Bast F (2014) Insulin receptor (IR) and insulin-like growth factor receptor 1 (IGF-1R) signaling systems: novel treatment strategies for cancer. Med Oncol 31(1):805 24. Zong CS, Zeng L, Jiang Y, Sadowski HB, Wang L-H (1998) Stat3 plays an important role in oncogenic Ros-and insulin-like growth factor I receptor-induced anchorage-independent growth. J Biol Chem 273(43):28065–28072 25. Resnicoff M, Baserga R (1998) The Role of the Insulin-like Growth Factor I Receptor in Transformation and Apoptosis. Ann N Y Acad Sci 842(1):76–81 26. Zeng H, Datta K, Neid M, Li J, Parangi S, Mukhopadhyay D (2003) Requirement of different signaling pathways mediated by insulin-like growth factor-I receptor for proliferation, invasion, and VPF/VEGF expression in a pancreatic carcinoma cell line. Biochem Biophys Res Commun 302(1):46–55 27. Ding X-Z, Fehsenfeld DM, Murphy LO, Permert J, Adrian TE (2000) Physiological concentrations of insulin augment pancreatic cancer cell proliferation and glucose utilization by activating MAP kinase, PI3 kinase and enhancing GLUT-1 expression. Pancreas 21 (3):310–320 28. Levitt RJ, Pollak M (2002) Insulin-like growth factor-I antagonizes the antiproliferative effects of cyclooxygenase-2 inhibitors on BxPC-3 pancreatic cancer cells. Cancer Res 62 (24):7372–7376 29. Hu H, Han T, Zhuo M, Wu LL, Yuan C, Wu L, Lei W, Jiao F, Wang L-W (2017) Elevated COX-2 expression promotes angiogenesis through EGFR/p38-MAPK/Sp1-dependent signalling in pancreatic cancer. Sci Rep 7(1):470 30. Wolpin BM, Michaud DS, Giovannucci EL, Schernhammer ES, Stampfer MJ, Manson JE, Cochrane BB, Rohan TE, Ma J, Pollak MN (2007) Circulating insulin-like growth factor binding protein-1 and the risk of pancreatic cancer. Cancer Res 67(16):7923–7928

128

G. Srivani et al.

31. Suzuki H, Li Y, Dong X, Hassan MM, Abbruzzese JL, Li D (2008) Effect of insulin-like growth factor gene polymorphisms alone or in interaction with diabetes on the risk of pancreatic cancer. Cancer Epidemiol Biomark Prev 17(12):3467–3473 32. Van Kruijsdijk RC, Van Der Wall E, Visseren FL (2009) Obesity and cancer: the role of dysfunctional adipose tissue. Cancer Epidemiol Biomark Prev 18(10):2569–2578 33. Ramos EJ, Xu Y, Romanova I, Middleton F, Chen C, Quinn R, Inui A, Das U, Meguid MM (2003) Is obesity an inflammatory disease? Surgery 134(2):329–335 34. Chen G, Goeddel DV (2002) TNF-R1 signaling: a beautiful pathway. Science 296 (5573):1634–1635 35. Parameswaran N, Patial S (2010) Tumor necrosis factor-α signaling in macrophages. Crit Rev Eukaryot Gene Expr 20(2):87–103 36. Aggarwal BB, Kunnumakkara AB, Harikumar KB, Gupta SR, Tharakan ST, Koca C, Dey S, Sung B: Signal transducer and activator of transcription-3, inflammation, and cancer: how intimate is the relationship? Ann N Y Acad Sci 2009, 1171(1):59–76 37. Heinrich PC, Behrmann I, Serge H, Hermanns HM, Müller-Newen G, Schaper F (2003) Principles of interleukin (IL)-6-type cytokine signalling and its regulation. Biochem J 374 (1):1–20 38. Hertzer KM, Xu M, Moro A, Dawson DW, Du L, Li G, Chang H-H, Stark AP, Jung X, Hines OJ (2016) Robust Early Inflammation of the Peri-pancreatic Visceral Adipose Tissue During Diet-Induced Obesity in the KrasG12D Model of Pancreatic Cancer. Pancreas 45(3):458 39. Kalaany NY, Sabatini DM (2009) Tumours with PI3K activation are resistant to dietary restriction. Nature 458(7239):725 40. Mattson MP, Allison DB, Fontana L, Harvie M, Longo VD, Malaisse WJ, Mosley M, Notterpek L, Ravussin E, Scheer FA (2014) Meal frequency and timing in health and disease. Proc Natl Acad Sci 111(47):16647–16653 41. Pollak M (2009) Do cancer cells care if their host is hungry? Cell Metab 9(5):401–403 42. Ogihara T, Asano T, Katagiri H, Sakoda H, Anai M, Shojima N, Ono H, Fujishiro M, Kushiyama A, Fukushima Y (2004) Oxidative stress induces insulin resistance by activating the nuclear factor-κB pathway and disrupting normal subcellular distribution of phosphatidylinositol 3-kinase. Diabetologia 47(5):794–805 43. Schieber M, Chandel NS (2014) ROS function in redox signaling and oxidative stress. Curr Biol 24(10):R453–R462 44. Martinez-Useros J, Li W, Cabeza-Morales M, Garcia-Foncillas J (2017) Oxidative stress: a new target for pancreatic cancer prognosis and treatment. J Clin Med 6(3):29 45. Pani G, Galeotti T, Chiarugi P (2010) Metastasis: cancer cell’s escape from oxidative stress. Cancer Metastasis Rev 29(2):351–378 46. Bedard K, Krause K-H (2007) The NOX family of ROS-generating NADPH oxidases: physiology and pathophysiology. Physiol Rev 87(1):245–313 47. Zorov DB, Juhaszova M, Sollott SJ (2014) Mitochondrial reactive oxygen species (ROS) and ROS-induced ROS release. Physiol Rev 94(3):909–950 48. Lee JK, Edderkaoui M, Truong P, Ohno I, Jang KT, Berti A, Pandol SJ, Gukovskaya AS (2007) NADPH oxidase promotes pancreatic cancer cell survival via inhibiting JAK2 dephosphorylation by tyrosine phosphatases. Gastroenterology 133(5):1637–1648 49. Vaquero EC, Edderkaoui M, Pandol SJ, Gukovsky I, Gukovskaya AS (2004) Reactive oxygen species produced by NAD (P) H oxidase inhibit apoptosis in pancreatic cancer cells. J Biol Chem 279(33):34643–34654 50. Narendhirakannan R, Hannah MAC (2013) Oxidative stress and skin cancer: an overview. Indian J Clin Biochem 28(2):110–115 51. Bastard J-P, Maachi M, Lagathu C, Kim MJ, Caron M, Vidal H, Capeau J, Feve B (2006) Recent advances in the relationship between obesity, inflammation, and insulin resistance. Eur Cytokine Netw 17(1):4–12

7

Diabetes with Pancreatic Ductal Adenocarcinoma

129

52. Zhang C, Cao S, Toole BP, Xu Y (2015) Cancer may be a pathway to cell survival under persistent hypoxia and elevated ROS: a model for solid-cancer initiation and early development. Int J Cancer 136(9):2001–2011 53. Ramanathan B, Jan K-Y, Chen C-H, Hour T-C, Yu H-J, Pu Y-S (2005) Resistance to paclitaxel is proportional to cellular total antioxidant capacity. Cancer Res 65(18):8455–8460 54. Fiorini C, Cordani M, Gotte G, Picone D, Donadelli M (2015) Onconase induces autophagy sensitizing pancreatic cancer cells to gemcitabine and activates Akt/mTOR pathway in a ROS-dependent manner. Biochim Biophys Acta 1853(3):549–560 55. Midaoui AE, Elimadi A, Wu L, Haddad PS, De Champlain J (2003) Lipoic acid prevents hypertension, hyperglycemia, and the increase in heart mitochondrial superoxide production. Am J Hypertens 16(3):173–179 56. Tirosh A, Potashnik R, Bashan N, Rudich A (1999) Oxidative stress disrupts insulin-induced cellular redistribution of insulin receptor substrate-1 and phosphatidylinositol 3-Kinase in 3T3-L1 adipocytes A Putative Cellular Mechanism For Impaired Protein Kinase B activation and glut4 translocation. J Biol Chem 274(15):10595–10602 57. Andersen DK, Andren-Sandberg Å, Duell EJ, Goggins M, Korc M, Petersen GM, Smith JP, Whitcomb DC (2013) Pancreatitis-diabetes-pancreatic cancer: summary of an NIDDK-NCI workshop. Pancreas 42(8):1227–1237 58. Duffy DL (2007) Genetic determinants of diabetes are similarly associated with other immunemediated diseases. Curr Opin Allergy Clin Immunol 7(6):468–474 59. Imamura M, Maeda S (2011) Genetics of type 2 diabetes: the GWAS era and future perspectives [Review]. Endocr J 58(9):723–739 60. Petersen GM, Amundadottir L, Fuchs CS, Kraft P, Stolzenberg-Solomon RZ, Jacobs KB, Arslan AA, Bueno-de-Mesquita HB, Gallinger S, Gross M et al (2010) A genome-wide association study identifies pancreatic cancer susceptibility loci on chromosomes 13q22.1, 1q32.1 and 5p15.33. Nat Genet 42(3):224–228 61. de Mendonça RL, Bouton D, Bertin B, Escriva H, Noël C, Vanacker J-M, Cornette J, Laudet V, Pierce RJ (2002) A functionally conserved member of the FTZ-F1 nuclear receptor family from Schistosoma mansoni. Eur J Biochem 269(22):5700–5711 62. Paré J-F, Malenfant D, Courtemanche C, Jacob-Wagner M, Roy S, Allard D, Bélanger L (2004) The fetoprotein transcription factor (FTF) gene is essential to embryogenesis and cholesterol homeostasis and is regulated by a DR4 element. J Biol Chem 279 (20):21206–21216 63. Repa JJ, Mangelsdorf DJ (1999) Nuclear receptor regulation of cholesterol and bile acid metabolism. Curr Opin Biotechnol 10(6):557–563 64. W-w L, Wang HW, Sum C, Liu D, Hew CL, Chung B (2000) Zebrafish ftz-f1 gene has two promoters, is alternatively spliced, and is expressed in digestive organs. Biochem J 348 (2):439–446 65. Benod C, Vinogradova MV, Jouravel N, Kim GE, Fletterick RJ, Sablin EP (2011) Nuclear receptor liver receptor homologue 1 (LRH-1) regulates pancreatic cancer cell growth and proliferation. Proc Natl Acad Sci 108:16927 66. Lin Q, Aihara A, Chung W, Li Y, Chen X, Huang Z, Weng S, Carlson RI, Nadolny C, Wands JR (2014) LRH1 promotes pancreatic cancer metastasis. Cancer Lett 350(1–2):15–24 67. Brissova M, Shiota M, Nicholson WE, Gannon M, Knobel SM, Piston DW, Wright CV, Powers AC (2002) Reduction in pancreatic transcription factor PDX-1 impairs glucosestimulated insulin secretion. J Biol Chem 277(13):11225–11232 68. Ashizawa S, Brunicardi FC, Wang X-P (2004) PDX-1 and the pancreas. Pancreas 28 (2):109–120 69. Annicotte J-S, Fayard E, Swift GH, Selander L, Edlund H, Tanaka T, Kodama T, Schoonjans K, Auwerx J (2003) Pancreatic-duodenal homeobox 1 regulates expression of liver receptor homolog 1 during pancreas development. Mol Cell Biol 23(19):6713–6724 70. Bell M, Crawford H (2006) The role of PDX-1 in the regulation of the MMP-7 gene expression in pancreatic cancer. In: AACR; 2006

130

G. Srivani et al.

71. Roy N, Takeuchi KK, Ruggeri JM, Bailey P, Chang D, Li J, Leonhardt L, Puri S, Hoffman MT, Gao S et al (2016) PDX1 dynamically regulates pancreatic ductal adenocarcinoma initiation and maintenance. Genes Dev 30(24):2669–2683 72. Defronzo RA (2009) Banting Lecture. From the triumvirate to the ominous octet: a new paradigm for the treatment of type 2 diabetes mellitus. Diabetes 58(4):773–795 73. Singh S, Singh PP, Singh AG, Murad MH, McWilliams RR, Chari ST (2013) Anti-diabetic medications and risk of pancreatic cancer in patients with diabetes mellitus: a systematic review and meta-analysis. Am J Gastroenterol 108(4):510 74. Landman GWD, Kleefstra N, van Hateren KJJ, Groenier KH, Gans ROB, Bilo HJG (2010) Metformin associated with lower cancer mortality in type 2 diabetes: ZODIAC-16. Diabetes Care 33(2):322–326 75. DeCensi A, Puntoni M, Goodwin P, Cazzaniga M, Gennari A, Bonanni B, Gandini S (2010) Metformin and cancer risk in diabetic patients: a systematic review and meta-analysis. Cancer Prev Res (Phila) 3(11):1451–1461. https://doi.org/10.1158/1940-6207.CAPR-10-0157 76. Currie C, Poole C, Gale E (2009) The influence of glucose-lowering therapies on cancer risk in type 2 diabetes. Diabetologia 52(9):1766–1777 77. Li D, Yeung S-CJ, Hassan MM, Konopleva M, Abbruzzese JL (2009) Antidiabetic therapies affect risk of pancreatic cancer. Gastroenterology 137(2):482–488 78. Li D (2012) Diabetes and pancreatic cancer. Mol Carcinog 51(1):64–74 79. Isoda K, Young JL, Zirlik A, MacFarlane LA, Tsuboi N, Gerdes N, Schonbeck U, Libby P (2006) Metformin inhibits proinflammatory responses and nuclear factor-κB in human vascular wall cells. Arterioscler Thromb Vasc Biol 26(3):611–617 80. Gallagher EJ, LeRoith D (2011) Diabetes, cancer, and metformin: connections of metabolism and cell proliferation. Ann N Y Acad Sci 1243(1):54–68 81. Viollet B, Guigas B, Sanz Garcia N, Leclerc J, Foretz M, Andreelli F (2012) Cellular and molecular mechanisms of metformin: an overview. Clin Sci (Lond) 122(6):253–270 82. Hadad SM, Fleming S, Thompson AM (2008) Targeting AMPK: a new therapeutic opportunity in breast cancer. Crit Rev Oncol Hematol 67(1):1–7 83. Shaw RJ, Lamia KA, Vasquez D, Koo S-H, Bardeesy N, Depinho RA, Montminy M, Cantley LC (2005) The kinase LKB1 mediates glucose homeostasis in liver and therapeutic effects of metformin. Science (New York, NY) 310(5754):1642–1646 84. Batandier C, Guigas B, Detaille D, El-Mir M, Fontaine E, Rigoulet M, Leverve XM (2006) The ROS production induced by a reverse-electron flux at respiratory-chain complex 1 is hampered by metformin. J Bioenerg Biomembr 38(1):33–42 85. He L, Wondisford FE (2015) Metformin action: concentrations matter. Cell Metab 21 (2):159–162 86. Gwinn DM, Shackelford DB, Egan DF, Mihaylova MM, Mery A, Vasquez DS, Turk BE, Shaw RJ (2008) AMPK phosphorylation of raptor mediates a metabolic checkpoint. Mol Cell 30(2):214–226 87. Williams T, Brenman JE (2008) LKB1 and AMPK in cell polarity and division. Trends Cell Biol 18(4):193–198 88. He L, Sabet A, Djedjos S, Miller R, Sun X, Hussain MA, Radovick S, Wondisford FE (2009) Metformin and insulin suppress hepatic gluconeogenesis through phosphorylation of CREB binding protein. Cell 137(4):635–646 89. Pearce EL, Walsh MC, Cejas PJ, Harms GM, Shen H, Wang L-S, Jones RG, Choi Y (2009) Enhancing CD8 T-cell memory by modulating fatty acid metabolism. Nature 460 (7251):103–107 90. Liang J, Shao SH, Xu Z-X, Hennessy B, Ding Z, Larrea M, Kondo S, Dumont DJ, Gutterman JU, Walker CL (2007) The energy sensing LKB1–AMPK pathway regulates p27 kip1 phosphorylation mediating the decision to enter autophagy or apoptosis. Nat Cell Biol 9 (2):218 91. Li X, Li T, Liu Z, Gou S, Wang C (2017) The effect of metformin on survival of patients with pancreatic cancer: a meta-analysis. Sci Rep 7(1):5825–5825

7

Diabetes with Pancreatic Ductal Adenocarcinoma

131

92. Kisfalvi K, Eibl G, Sinnett-Smith J, Rozengurt E (2009) Metformin disrupts crosstalk between G protein-coupled receptor and insulin receptor signaling systems and inhibits pancreatic cancer growth. Cancer Res 69(16):6539–6545 93. Sadeghi N, Abbruzzese JL, Yeung S-CJ, Hassan M, Li D (2012) Metformin use is associated with better survival of diabetic patients with pancreatic cancer. Clin Cancer Res 18 (10):2905–2912 94. Schneider MB, Matsuzaki H, Haorah J, Ulrich A, Standop J, Ding XZ, Adrian TE, Pour PM (2001) Prevention of pancreatic cancer induction in hamsters by metformin. Gastroenterology 120(5):1263–1270 95. Krisztina K, Aune M, James S-S, Guido E, Enrique R (2013) Metformin inhibits the growth of human pancreatic cancer xenografts. Pancreas 42(5):781 96. Tan X-L, Bhattacharyya KK, Dutta SK, Bamlet WR, Rabe KG, Wang E, Smyrk TC, Oberg AL, Petersen GM, Mukhopadhyay D (2015) Metformin suppresses pancreatic tumor growth with inhibition of NFκB/STAT3 inflammatory signaling. Pancreas 44(4):636–647 97. Qian W, Li J, Chen K, Jiang Z, Cheng L, Zhou C, Yan B, Cao J, Ma Q, Duan W (2018) Metformin suppresses tumor angiogenesis and enhances the chemosensitivity of gemcitabine in a genetically engineered mouse model of pancreatic cancer. Life Sci 208:253–261 98. Sola D, Rossi L, Schianca GPC, Maffioli P, Bigliocca M, Mella R, Corlianò F, Fra GP, Bartoli E, Derosa G (2015) Sulfonylureas and their use in clinical practice. Arch Med Sci 11 (4):840 99. Ashcroft FM (1996) Mechanisms of the glycaemic effects of sulfonylureas. Horm Metab Res 28(09):456–463 100. Bowker SL, Majumdar SR, Veugelers P, Johnson JA (2006) Increased cancer-related mortality for patients with type 2 diabetes who use sulfonylureas or insulin. Diabetes Care 29 (2):254–258 101. Monami M, Lamanna C, Balzi D, Marchionni N, Mannucci E (2009) Sulphonylureas and cancer: a case–control study. Acta Diabetol 46(4):279 102. Smith U, Gale EAM (2009) Does diabetes therapy influence the risk of cancer? Diabetologia 52(9):1699–1708 103. Gerstein HC (2010) Does insulin therapy promote, reduce, or have a neutral effect on cancers? JAMA 303(5):446–447 104. Wang F, Gupta S, Holly EA (2006) Diabetes mellitus and pancreatic cancer in a populationbased case-control study in the San Francisco Bay Area, California. Cancer Epidemiol Biomark Prev 15(8):1458–1463 105. Andersen DK, Korc M, Petersen GM, Eibl G, Li D, Rickels MR, Chari ST, Abbruzzese JL (2017) Diabetes, pancreatogenic diabetes, and pancreatic cancer. Diabetes 66(5):1103–1110

8

Role of Inflammatory Cytokines in the Initiation and Progression of Pancreatic Cancer Madanraj Appiya Santharam and Vignesh Dhandapani

Abstract

It is essential to identify different targets for pancreatic cancer (PC) as it is asymptomatic until it has metastasized to other organs. After which time, the convention methods to control cancer such as surgery and chemotherapy is no longer an option. Over the years, it has been noted that the inflammation is a major cause of initiation of pancreatitis, leading to cancer. Various inflammatory cytokines initiate and play a role in progression of this cancer. In this chapter, we will be discussing about major inflammatory cytokines that have been identified over the years. For each cytokine, we will be looking at the source of the cytokines, the signaling mechanism of that cytokine, its role in other cancers followed by pancreatic cancer. Since the survival of many diagnosed PC patients are limited to months, understanding the origins and basic cascades of these different cytokines could better help in identifying the targets and developing different therapeutic approaches in screening of the patients with this disease. Keywords

Cytokines · Pancreatic Cancer · Inflammation · Extra-cellular Matrix · Interleukins

Abbreviations AP-1 APC

Activating Protein 1 Antigen Presenting Cells

M. A. Santharam (*) University of Leicester, Leicester, UK e-mail: [email protected] V. Dhandapani Technische Universität Dresden, Dresden, Germany # Springer Nature Singapore Pte Ltd. 2019 G. P. Nagaraju, A. BM Reddy (eds.), Exploring Pancreatic Metabolism and Malignancy, https://doi.org/10.1007/978-981-32-9393-9_8

133

134

ASK BAD BMP CAF CD CLC CNTF DC ECM EMT ERK FADD GDF GDNF GM-CSF IFN IKK IL JAK JNK LAP LIF LLC LTBP MAPK/MEK MDSC MHC MLK MMP NF-κB NK OSM PanIN PDAC PI3K PSC PTEN RAGE RIP ROS RTK SBE SCID SMA SMAD

M. A. Santharam and V. Dhandapani

Apoptosis Signal-regulating Kinase BCL2 Associated Death Promoter Bone Morphogenetic Protein Cancer Associated Fibroblasts Cluster of Differentiation Cardiotrophin-Like Cytokine Ciliary Neurotrophic Factor Dendritic Cells Extra Cellular Matrix Epithelial-Mesenchymal Transition Extracellular Signal-Regulated Kinases Fas-Associated protein with Death Domain Growth and Differentiation Factors Glial cell line-Derived Neurotrophic Factor Granulocyte Macrophage Colony Stimulating Factor Interferon IκB Kinase Interleukin Janus Kinase c-Jun N-terminal Kinase Latency-Associated Peptide Leukemia Inhibitory Factory Large Latent Complex Latent TGF-β Binding Proteins Mitogen Activated Protein Kinases Myeloid Derived Suppressor Cells Major Histocompatibility Complex Mixed-Lineage Protein Kinases Matrix Metalloproteinase Nuclear Factor-κB Natural Killer Oncostatin M Pancreatic Intraepithelial Neoplasia Pancreatic Ductal Adenocarcinoma Phosphoinositide 3-Kinase Pancreatic Stellate Cells Phosphatase and Tensin Homolog Receptor for Advanced Glycation End products Receptor Interacting Protein Reactive Oxygen Species Receptor Tyrosine Kinase STAT Binding Elements Severe Combined Immunodeficiency Smooth Muscle Actin Small Mothers Against Decapentaplegic

8

Role of Inflammatory Cytokines in the Initiation and Progression of. . .

SOCS SODD SRF STAT TAK TGF TME TNF TRADD TRAF Tregs Tyk VEGF

8.1

135

Suppressor of Cytokine Signaling Silencer of Death Domain protein Serum Response Factor Signal Transducer and Activator of Transcription TGF-β-Activated Kinase Transforming Growth Factor Tumor Microenvironment Tumor Necrosis Factor TNF Receptor-Associated Death Domain protein TNF Receptor-Associated Factor Regulatory T-cells Tyrosine Kinase Vascular Endothelial Growth Factor

Introduction

Pancreatic Ductal Adenocarcinoma (PDAC) being asymptomatic during its progression makes it very lethal with the 5-year survival rate as low as 6% [147], as conventional methods such as surgery or chemotherapy would not be beneficial due to the advanced stages of the disease. PDAC is second, just behind colon cancer, in estimated new cases and mortalities in the US for cancers originating in the digestive system [147]. Familial history increases the risk of getting mutation in oncogenes, contributing to the increase in number of new patients [15, 87]. Recent developments in the immunotherapy has made people develop multiple clinical trials involving the PD-1/CTLA-4 combination [34, 48, 78]. The limitations of these trials could stem from the ability of the cancer to maintain an immunosuppressive tumor microenvironment [41]. The tumor microenvironment (TME) of PDAC has an extensive layer of stroma (known as desmoplasia) due to the presence of cancer associated fibroblasts (CAFs) [159] and a large number of immunosuppressive cells that is consisted mainly of γδT-cells, regulatory T-cells (T-regs), myeloid derived suppressor cells (MDSCs) and M2 macrophages [32, 37, 76, 100]. Anti-inflammatory cytokines secreted these inflammatory infiltrates play a major role in aiding the development of PDAC. Chronic pancreatitis is also an important initiator of PDAC. Chronic inflammation in the pancreas causes repeated injury to the parenchymal cells resulting in a fibrogenic response [123]. Induced by pro-inflammatory cytokines such as IL-1β, IL-6, TNF-α [89], this inflammation causes the quiescent pancreatic stellate cells to get activated and obtain a myofibroblast-like phenotype. These myofibroblasts secrete pro-fibrogenic cytokine TGF-β and other growth factors. This further induces the fibrosis in extracellular matrix (ECM) and ultimately leading to pancreatic cancer [106]. In this chapter, we will be expanding on the various cytokines – antiinflammatory and pro-inflammatory cytokines – secreted by different cell types and how they affect the initiation and progression of PDAC.

136

8.2

M. A. Santharam and V. Dhandapani

Cytokines in Pancreatic Cancer

Cytokines are small, low-molecular weight proteins secreted by tumor cells and surrounding cells into the TME that modulate the biological metabolic processes such as differentiation, proliferation and migration of various cells [40, 117]. Inflammatory cytokines, both pro- and anti-inflammatory cytokines, have a clinical significance as they are upregulated in pancreatic cancer and their expression has a negative effect on the clinical outcome [43, 49]. There is an interplay between the T-regs, macrophages and the tumor cells in secreting anti-inflammatory cytokines to facilitate tumor growth and invasion. Furthermore, angiogenesis and immune evasion are supported by pro-inflammatory cytokines and growth factors [13, 56, 135].

8.3

Anti-inflammatory Cytokines

8.3.1

Interleukin-10 (IL-10)

IL-10 signals through Janus Kinase (JAK) – Signal Transducer and Activator of Transcription (STAT) signaling cascade as the other members of its Interleukin family. It is the founding member of the IL-10 interleukin superfamily which also consists of Interleukin 19 (IL-19), Interleukin 20 (IL-20), Interleukin 22 (IL-22), Interleukin 24 (IL-24), Interleukin 26 (IL-26) and type III Interferons (IFNs) [17]. IL-10 is secreted by a host of immune and non-immune cells based on the type of infection or inflammation [80, 116]. The soluble IL-10 binds to the extracellular tetrameric IL-10 receptor (IL-10R) consisting of two subunits – IL-10R1 and IL-10R2, that homodimerize to initiate the signaling. The dimerized receptors activate the intracellular Tyrosine Kinase-2 (Tyk2) associated with the IL-10R1 subunit that simultaneously activates the JAK1 constitutively bound on the IL-10R2 subunit. This causes STAT3, a nuclear translocating transcription factor to induce gene expression of anti-inflammatory genes, to dock on the IL-10R components and to get phosphorylated by JAK1 and Tyk2 [164]. The phosphorylated STAT3 un-docks and dimerizes in the cytoplasm before entering into the nucleus and binding to the STAT Binding Elements (SBE), a promoter for IL-10 responsive genes, to initiate the transcription of IL-10 target genes [144] (Fig. 8.1). Suppressor of Cytokine Signaling 3 (SOCS3), a part of the 8-member SOCS family of proteins, is one of the downstream target genes is induced by IL-10 [22] that acts a negative feedback regulator of its inducer [16]. Immediately after that publication, Ding et al. in 2003, showed that another member of that family, SOCS1 is also induced by IL-10 and it controls the IL-10-mediated immune responses [39]. IL-10 functions as a prominent cytokine in regulating the inflammation in our body. Its major role is to regulate the immune cells to prevent them from causing acute tissue damage during an infection [31]. The source of IL-10 was primarily identified to be Th2 helper CD4+ T-cells in 1989 [51] but was later linked to other CD4+ T-cell types such as Th1 CD4+ T-cells and CD4+CD25+FoxP3+ T-regs, the dendritic cells (DCs) and macrophages to inhibit the activity of NK-cells, B-cells,

8

Role of Inflammatory Cytokines in the Initiation and Progression of. . .

137

Fig. 8.1 The different cascades followed by the inflammatory cytokines leading to PDAC. The pro-inflammatory cytokines activate the signaling pathways to aid in the EMT of the cancer cells and the anti-inflammatory cells facilitate this cell proliferation through development of fibrous tissue in the tumor microenvironment surrounding the tumor. EMT epithelial mesenchymal transition, IL Interleukin, TNF-α Tumor Necrosis Factor alpha, TGF-β Transforming Growth Factor beta, PSC Pancreatic Stellate Cells

CD8 T-cells and the other innate immune cells [2, 156]. This is effectively done indirectly by down-regulating the Major Histocompatibility Complex (MHC-II) expression and its co-stimulatory molecule CD80/CD86 on the antigen presenting cells (APCs) such as DCs and Macrophages [81]. IL-10 also inhibits the production of pro-inflammatory cytokines such as IL-1β, IL-5, IL-6, IL-12, TNF-α [52, 139] and other chemokines [111] by DCs and Macrophages. The APCs are also directly targeted by IL-10 through its autocrine secretion which causes inhibition of DC trafficking to lymph nodes to present the antigens to the naïve T-cells to differentiate into activated T-cells [33]. Recently, it has been identified that IL-10 production is enhanced by IL-4 secreted by Th2 cells influencing the Th1 cells towards a selfregulatory phenotype [108]. Therefore, during any infection IL-10 uses direct and indirect methods to inhibit the activation of cytotoxic cells. A context dependent, paradoxical role has been observed for IL-10 in cancer. In some cancers, it acts a tumor promoter by promoting the tumor immune evasion whereas in some conditions, it acts a tumor suppressor by inhibiting the tumor growth through increased IL-10 secretion [99]. A study showed reduced tumor growth by using mouse models injected with cells overexpressing IL-10 [61] and it was further corroborated by the requirement of IL-10 for enabling the memory of CD8 T-cells [53]. In pancreatic cancer, IL-10 acts as a tumor promoter as high levels of IL-10 was measured in serum of PDAC clinical patients, who were associated with a poor survival rate [47]. Other previous studies have also observed increased level of IL-10 in serum along with high expression levels of IL-1, IL-6, IL-8, IL-17,

138

M. A. Santharam and V. Dhandapani

IL-23 and TNF-α [19, 157]. Mesothelin, a surface marker associated with tumor growth in multiple neoplasms including pancreatic cancer [26], is recognized better by mesothelin specific CD8 T-cells following inhibition of IL-10 in the pancreatic tumor microenvironment [12]. Mechanistically, pancreatic tumor progression is also due to the development of an immunosuppressive environment created by the infiltration of IL-10 secreting γδT-cells and blocking the efficacy of αβT-cells in the PDAC [110]. In addition to this, in vitro pancreatic cancer cells having a KRASG12D mutation converts CD4+CD25- T-cells to T-regs and activate the MEK/ERK pathway resulting in the secretion of IL-10 and TGF-β [27]. Activation of MAPK kinase pathway, especially ERK1 and ERK2, is essential for the induction of IL-10 by Th1 cells [138]. All these data indicate an essential role of IL-10 in prognosis and pathogenesis of PDAC.

8.4

Transforming Growth Factor (TGF-b)

TGF-β was initially identified in early 1980s, along with TGF-α, as a protein which had the capability to form large colonies of murine sarcoma virus-transformed 3T3 cells in soft agar [5]. When it was first discovered, the function of TGF-β was widely conflicted as it had more than one function and this was against the notion at that time that each hormone had only one function in the body. But eventually it was identified that TGF-β has a context dependent signaling in cells [102]. TGF-β superfamily consists glial cell line-derived neurotrophic factor (GDNF), bone morphogenetic protein (BMP)/growth and differentiation factors (GDF), Activins and Inhibins, and TGF-β ligands [7, 137]. The polypeptide TGF-β ligands are constituted by TGFβ1, TGFβ2 and TGFβ3, that bind to its receptors TGFβR1 and TGFβR2, with the latter being the specific receptor required for all the three ligands [4, 130]. The production of latent inactive TGFβ is started in the Endoplasmic Reticulum (ER) consisting of the latency-associated peptide (LAP) region and TGF-β region as pre-pro-TGF-β peptide [119]. This peptide dimerizes as the low-glycosylated pro-TGF-β and enters into the Golgi complex. In the Golgi, furin binds to this 82-kDa protein and forms a high-glycosylated pro-TGF-β complex [119]. This complex could be immaturely secreted, or a more stable complex could form as high-glycosylated latent TGF-β complex that houses the TGF-β portion in between the LAP regions. This complex is secreted out as a latent TGF-β [119] . It could also be surface bound on T-regs [112]. The combination of latent TGF-β and LAP is called as the small latent complex and together with the Latent TGF-β binding proteins (LTBP 1-4) in the Extra-Cellular Matrix (ECM), it is called as large latent complex (LLC) [165]. The integrins of the cells bind to the LAP of the LLC and pull the LAP to open up the TGF-β for its receptor. The ECM and MatrixMetalloproteinases (MMP) break the TGF-β from the LLC. If the ECM is rigid, it is easier for the cytokine to get released and then MMP2 and MMP9 cleave the LAP to release TGF-β [165]. Once the active TGF-β binds to its receptors, they signal through the intracellular Serine/Threonine Kinase (Ser/Thr) domain unlike the other cytokines which signal

8

Role of Inflammatory Cytokines in the Initiation and Progression of. . .

139

via the Tyrosine Kinase domain (RTK). The TGFβR2 phosphorylates TGFβR1 of the receptor complex and classical TGFβ signaling pathway involves recruitment of the SMAD2 and SMAD3 and subsequently getting phosphorylated themselves [121]. These activated SMADs bind to the SMAD4 before translocating to nucleus and binding to the promoter regions of its target genes along with some of the other transcription factors to initiate various cellular responses [75] (Fig. 8.1). The SMAD7, co-binding with NEDD4 and Smurf1/2, is the natural regulator of the pathway [1]. The non-canonical pathway activates numerous other pathways such as JNK, P38, PI3K/AKT pathways [75, 121]. The primary function of TGF-β is to be an anti-inflammatory cytokine and suppress pro-inflammatory responses of many immune cells. The variety of immune cells regulated by TGF-β has been extensively reviewed by Mo et al. [92]. But in cancer, its role is paradoxical as it behaves as both tumor promoter and tumor suppressor in a context dependent manner. When the tumor is in its nascent stage, TGF-β tries to inhibit the growth of tumor by exerting growth arrest, causing apoptosis and preventing inflammation [67]. Initial studies also suggested the same as indicated by the absence of a functional TGF-β signaling in many human PDAC cell lines [158]. But as the tumor progresses, it turns detrimental to the body by promoting cancer through induction of epithelial-mesenchymal transition (EMT), promoting immune evasion and angiogenesis. Cancer stem cells enriched from the adjacent population of a pancreatic cell line showed enhanced ability to change from epithelial morphology to mesenchymal-like morphology on addition of TGF-β [82]. It also has pro-metastatic abilities where it modifies the TME to facilitate the extravasation of tumor cells and promotes colonization [67]. In 2003, it was shown that canonical TGF-β signaling was defective in pancreatic cancer cell lines [114]. Moreover, clinical studies have determined high levels of TGF-β in PDAC patients [77, 171]. This high level of TGF-β in the plasma heavily promotes fibrosis in the chronic pancreatitis through the increase of pancreatic stellate cells (PSCs) and induction of MMPs in ECM [46, 145]. PDAC cells secrete TGF-β to transform quiescent PSC into activated PSC that have an autocrine secretion of various cytokines including TGF-β and this activated PSC increases matrix deposition, upregulates MMPs and downregulates TIMPs ultimately resulting in fibrosis in chronic pancreatitis and cancer [21]. A cascade of MAPK signaling pathways – JNK, p38 and ERK – is activated in PSCs by TGF-β to induce fibrosis. Addition of pharma logical inhibitors of these pathways downregulated the expression of TGF-β induced fibronectin and α-SMA, thereby fibrosis [169]. Recently, it has been reported that pancreatic cancer cell-derived exosomes could induce the fibrogenic proteins mentioned above when they were added to human primary PSCs. This study identified that TGF-β was an upstream target of these cancer derived exosomes [101]. To the pancreatic cancer themselves, TGF-β downregulated a potent tumor suppressor PTEN by activating NF-κB [28]. Canonical Wnt signaling is also induced by TGF-β through production of homeobox transcription factor CUTL1, whose target gene is Wnt-5A ligand thereby promoting pancreatic cancer [131]. Another transcription factor KLF11 was inhibited by oncogenic ERK/MAPK

140

M. A. Santharam and V. Dhandapani

pathway in PDAC, which had the function of terminating the inhibitory SMAD7 in the canonical TGF-β signaling in normal epithelial cells [45]. Therapeutic inhibitors at various levels of TGF-β signaling pathway in different cancers are being researched all over the world. Specifically, for the pancreatic carcinoma, E-cadherin is downregulated in cell lines of pancreatic cancer [151] but TGFβR2 targeting miR-655 has been identified a potential inhibitor which targets the EMT by inducing the expression of E-cadherin using Panc-1 cell line [63]. A kinase inhibitor, LY2109761, targeting the TGFβR1 and TGFβR2 significantly suppressed metastasis in orthotropic PC [107]. At the ligand level, BMP2 – a ligand from the TGF-β superfamily – inhibits the activation of PSCs by controlling the induction of α-SMA, fibronectin and collagen Ia [58]. From all these studies, it is evident that TGF-β is a key cytokine in promoting pancreatic cancer and targeting of its defective pathway in cancer cells would lead us in finding a solution for reduction of fibrosis in PC.

8.5

Pro-inflammatory Cytokines

8.5.1

Tumor Necrosis Factor-a (TNF-a)

TNF-α is from the TNF super family proteins that includes type II transmembrane proteins consisting of TNF homology domains and form trimers [8]. When the TNF-α is initially translated, it is membrane bound. It is in its inactive form, similar to TGF-β, is of size 26-28-kDa and is predominantly produced by macrophages and CD4+ lymphocytes, but also by eosinophils, mast cells, neurons, neutrophils and NK cells [57, 120]. Inactive TNF-α is converted to a soluble form (sTNF) by the metalloprotease called as TNF alpha converting enzyme [18]. The secreted TNF-α assumes a pyramid shape and has a smaller molecular weight of 17-kDa. Both the types of the cytokine, soluble and membrane bound TNF-α, contributes to distinct biological functions [122]. TNF-α can bind either to a 55-kDa receptor 1 (TNFR1) or a 75-kDa receptor 2 (TNFR2) [155]. The binding of the ligand to its receptors leads the receptors to trimerize. This subsequently enables the release of inhibitory protein silencer of death domain protein (SODD) and hence the adaptor protein TNF receptorassociated death domain protein (TRADD) binds to the death domain leading to the activation of three pathways [24, 160]. The first of the pathways is the activation of Nuclear Factor-κB (NF-κB) signaling cascade. NF-κB, as a transcription factor, enables the transcription of variety of proteins that aid in cell survival, proliferation and inflammatory response [60]. TRADD aids in the binding of TNF receptorassociated factor 2 (TRAF2) and Receptor Interacting Protein (RIP). TRAF2 recruits IκB kinase (IKK) to bind to itself. IKK phosphorylates the inhibitory protein IκBα, that is found to be bound to NF-κB complex. and releases NF-κB to be translocated to the nucleus for its transcriptional activity [65] (Fig. 8.1). The second pathway is the activation of MAP kinase pathway. JNK (c-Jun N-terminal kinases) are kinases that phosphorylates specific transcription domains such as c-Jun. TNF-α is found to

8

Role of Inflammatory Cytokines in the Initiation and Progression of. . .

141

be strongly activating the MAPK pathway associated with JNK phosphorylation. TRAF2 and Rac complex enables the activation of kinases of mixed-lineage protein kinases (MLK2/MLK3), MEK kinase 1 (MEKK1), TGF-β-activated kinase (TAK1), Apoptosis signal-regulating kinase 1 (ASK1). These kinases phosphorylate MAPK kinase 7 (MKK7) and in-turn activate JNK kinases [84]. The third pathway is the induction of death signaling. TNF-α-induced cell death consumes a very small portion compared to the inflammatory signaling. TRADD recruits Fas-associated protein with death domain (FADD) and this consequently binds to cysteine protease Caspase-8. Higher concentration of caspase 8 results in proteolytic activation of other caspases and results in cell apoptosis [59]. In the sequence of the pathways, there has been conflicting pathways reported to take place due to the cross talk between the components involved. These cross talks could enable differential expression of the transcription factors and thereby form a critical part in shaping the biological functions [118]. The amount of reactive oxygen species (ROS) is also found to influence and shift the balance favoring one of the three pathways [154]. TNFR1 is responsible for causing inflammation by enabling the activation of transcription factors in the NF-κB, JNK and MAPK signaling pathways [36]. Activated TNFR1 induce cell death by many signaling pathways including activation of pro-apoptotic Bcl-2 family of proteins and ROS [38]. TNFR2 is found to induce an anti-inflammatory response [44]. TNF-α which is produced by the tumor cells influences the surrounding cells presenting in the TME to produce a cascade of chemokines and other cytokines. This results in the promotion of tumor growth, angiogenesis, metastases and finally forms the immune evasive environment [9, 143, 161]. The mechanism of TNF-α causing pancreatic cancer is an increase of pro-inflammatory macrophages (M1) as compared to the anti-inflammatory macrophages (M2). With the increase in M1, there is an increase in the secretion of TNF-α. This in-turn induces DNA damage due to the upregulation in the production of reactive nitrogen species (RNS) and ROS [9]. The altered mechanism in the DNA repair fails to expand normal cells while enhancing the tumor cell expansion. In the TME, the PDAC tumor cells and the M1 macrophages produce TNF-α and this TNF-α induces the proliferation of the tumor cells and M2 macrophages and hence produces stroma and desmoplasia [9, 89]. Finally, this mechanism results in a condition called cachexia and insulin resistance. In 1993, studies indicated that use of anti-TNF-α antibodies could inhibit the progression of cancer [9]. The hypothesis of these antibodies impairing the desmoplastic tumor stroma was verified through in vitro mice experiments as there was a decrease in the number of cellular components and the amount of collagen [181]. Food and Drug Administration (FDA) and European Medicines Agency (EMA) have approved five potent anti-TNF-α biologics – adalimumab, certolizumab, etanercept, infliximab, and golimumab. In vivo studies comparing infliximab and etanercept proved that the former elicits higher anti-tumor capacity [44]. The anti-TNF-α treatment has resulted in better effect in a short term, while in long term, recurrence of PDAC or an onset have also been reported in rare cases [6, 96, 163]. PDAC has been characterized by the upregulation of oncogenic miRNA

142

M. A. Santharam and V. Dhandapani

and downregulation of tumor suppressing miRNA [64]. Studies with patients with PDAC and other diseases treated with anti-TNF-α indicated that certain miRNA such as miR-197, miR-23a, miR-221, miR-227, let-7d were found to be upregulated and associated with pancreatic cancer [23, 29, 35, 54, 88, 124, 172]. Many more therapeutic targets are still under development for PDAC.

8.5.2

Interleukin-6 (IL-6)

IL-6 is a pleiotropic cytokine that has dual inflammatory properties. It is secreted by the osteoblasts (to induce osteoclast formation) and the smooth muscle cells. It is also a potent inducer of acute response and aids in the differentiation of β-cells to Ig-secreting cells [152]. IL-6 belongs to family of cytokines that includes leukemia inhibitory factor (LIF), cardiotrophin-like cytokine (CLC), oncostatin M (OSM), IL-11 and ciliary neurotrophic factor (CNTF). Its structural form is a four-helix bundle [150]. The signaling pathway for IL-6 occurs via the signal transducing complex gp-130 (CD130). Upon binding to the IL-6 cytokine, the IL-6 receptor (IL-6R) and the gp-130 form a complex and thereby activate the signal transduction through JAKSTAT [66]. IL-6 receptor has been identified to be of two types – soluble (sIL-6R) and membrane bound IL-6R (mIL-6R). IL-6 either binds to the soluble IL-6 receptor (sIL-6R) to follow the trans signaling pathway [115] or to the membrane bound (mIL-6R), following the classical signaling pathway [42, 133, 134]. mIL-6R is expressed only by certain types of cells such as the hepatocytes, macrophages and neutrophils. sIL-6R has been identified to be a result of ectodomain shedding from the cell and alternate splicing. The sIL-6R interaction with IL-6 is described to cause a proinflammatory response [132, 140]. IL-6 binds to the receptor and activates JAK and Ras mediated signaling which leads to the phosphorylation of STATs (STAT2 and STAT3) and SHP2 domain containing tyrosine phosphate [113]. Activated STAT3 forms dimer and translocate to the nucleus to activate genes of the STAT3 response elements (Fig. 8.1). SHP2 connects the receptor to Ras/MAP kinase pathway, which is essential for mitogenic activity [93, 141]. The Ras mediated pathway activates the proteins downstream of MAP kinases and activates transcription factors such as NF-κB and Elk1. These transcription factors in combination with other transcription factors such as SRF (Serum Response Factor) and Activating Protein-1 (AP-1) further regulate a set of promotors [11]. IL-6 also activates the Phosphoinositide-3 kinase pathway (PI3K/ Akt/NF-κB cascade) which results in the maximum anti-apoptotic effect of IL-6 against TGF-β. This anti-apoptotic effect is due to the phosphorylation of BCL2 associated death promoter (BAD) by Akt. Termination of IL-6 signaling happens through tyrosine phosphatases, proteases and JAK kinase inhibitors (SOCS, PIAS) [170]. IL-6 associated with PDAC has been identified due to the dysregulation of JAK2STAT3 signaling pathway activation [71]. STAT3 is an anti-apoptotic regulator of a varied number of genes responsible for angiogenesis, tumor growth and survival

8

Role of Inflammatory Cytokines in the Initiation and Progression of. . .

143

[176]. STAT3 dysregulation is found to be a key in cancer progression. A role of STAT3 in pancreatic lesion development and pancreatic intraepithelial neoplasia (PanIN) has been identified [30, 55]. In addition, IL-6 was also identified to be involved in the pancreatic cancer cell migration by the activation of a specific GTPase CDC42 [128]. Absence of IL-6 in lesions of PanIN’s resulted in the decreased activation of MAPK, Akt and STAT3 signaling mechanisms and hence had a reduced proliferation rate [179]. A link between pancreatic cancer development and inflammation has been long established [97, 109, 123]. IL-6 was examined to be over expressed in tissues of pancreatic cancer compared to the controls [14]. The serum level of the IL-6 is used as a biomarker of malignancy but the specificity and sensitivity of this test are highly variable [85]. Mesothelin, a pancreatic cancer marker, has been found to influence the IL-6 expression in PDAC [79]. Second molecule which influences this protein’s expression is the receptor for advanced glycation end products (RAGE) [24]. KRAS plays an important role in pancreatic tumorigenesis [83]. Zinc transporter ZIP4 is also associated with pancreatic cancer progression [178]. IL-6 expression was also observed to be increased in cases of hypoxia induced by miR-21, a pro-tumorigenic miRNA [10, 50]. IL-6 promotes angiogenesis and invasion by increasing the expression of Vascular Endothelial Growth Factor (VEGF) and Matrix metalloproteinase-2 (MMP-2) [72, 103, 153]. IL-6 enhances the expression of IL-5, IL-7, IL10, IL-13 and granulocyte macrophage-colony stimulating factor (GM-CSF), thereby providing a pro-tumorigenic environment in pancreatic cancer [50]. IL-6 increased the expression of neuropilin-1, which subsequently increased the MAPK signaling, thereby inducing chemo-resistance and anoikis in pancreatic cancer [50, 162]. KRASG12D model system was found to be the ideal model system to understand the initiation and development of PDAC [69]. IL-6 was also found to predominantly expressed in the infiltrating immune cells (F4/80+ macrophages) as compared to the acinar cells [91]. Chimeric anti-IL-6 monoclonal antibody called siltuximab was able to stabilize disease in ovarian cancer [173]. A humanized antibody, Tocilizumab, has successfully demonstrated to treat cancer cachexia and suppressed the growth of human squamous cell carcinoma [3, 146]. Unfortunately, no clinical trials identifying antiIL-6 for treatment of PDAC has been performed. Treating KPCY mice having PanIN with anti-IL-6 was able to reduce the count of precursor lesions [179]. Adjuvant and palliative inhibition of classical signaling by tocilizumab and trans signaling by sgp130Fc reduced tumor growth of pancreatic tumor cells in SCID mice [71]. AntiIL-6R antibodies targeted LY6Chi monocytes in mouse were able to inhibit the STAT3 activation and decreased tumor cell progression in vivo. IL-6R blockade coupled with chemotherapy was able to induce tumor apoptosis, regression and increased the survival [95]. Bazedoxifene was found to inhibit JAK1 binding and STAT3 phosphorylation. It also inhibited cell proliferation, viability and glycolysis of pancreatic cancer cells [25]. Suppression of IL-6 by sh-RNA increased cell apoptosis and increased the antitumor effect of gemcitabine [168]. Promising studies on the therapeutic targets for IL-6 in PC has been done in mouse models but clinical translations are yet to come.

144

8.5.3

M. A. Santharam and V. Dhandapani

Interleukin-17 (IL-17)

IL-17, a pro-inflammatory cytokine, is homo-dimeric and scereted by a CD4+ T helper cell subgroup called as T helper 17 cells (Th17 cells) in response to stimulation with IL-23 [70]. IL-17 is also produced by other adaptive immune cells such as γδ T-cells [62]. IL-17 can be further classified as six different proteins, from IL-17A to IL-17F, based on its amino acid sequence homology [74, 90, 136, 175]. IL-17A and IL-17F signal through the heterodimeric IL-17RA/RC receptors. The effect of IL-17A signaling is higher than the IL-17F signaling [182]. The IL-17R receptor complex undergoes a conformational change when the cytokine binds. This conformational change aids in homotypic interactions between the SEF/IL-17R domain (SEFIR) and signal adaptor Act1 [125]. This could cause the initiation of the canonical pathway, wherein the E3 Ubiquitin (Ub) ligase activity of Act1 could aid in the Lys-63 mediated ubiquitylation of the TRAF6 [142]. Furthermore, the canonical NF-κB, c/EBPβδ and MAPK pathways are activated [174]. This enables transcriptional activation of genes involved in the release of chemokines, proinflammatory cytokines and antimicrobial peptides. Non-canonical pathway is initiated by the phosphorylation of Act1 at the amino acid residue 311 by the IκB kinase (IKKi) and TBK-1 [20, 126]. TRAF2 and TRAF5 are recruited to the site of receptor complex and could sequester the effect of the RNA destabilizing factor ASF/SF2. On the contrary, mRNA stabilizing factor, HuR, is recruited that increases the half-life of mRNAs [68, 149]. IL-17E is found to signal through a heteromeric receptor, similar to 17A or 17F but the heterodimeric complex consists of IL-17RB, instead of IL-17RC, to bind to IL-17RA [129]. Similar to 17A canonical signaling cascade, intracellular portion of IL-17RA/RB complex also enables Act1 to recruit TRAF6, thereby activating the MAPK ad NF-κB pathways [98]. However, in contrast, Act1 could recruit TRAF4 aiding in further recruitment of E3 Ub ligase SMURF2. Ubiquitylation and degradation of IL-17RB inhibitor DAZAP2 helps in the subsequent IL-17E mediated signaling [177]. Act 1 independent activation of STAT5 could also initiate Th2 response [166]. The downstream signaling of IL-17B, C and D remains unclear. IL-17B is found to engage in proinflammatory secretion to enhance inflammation in pancreatic and breast cancer [73]. IL-17C is identified to signal through a heterodimeric IL-17RA/E complex [127] and activates the MAPK and NF-κB pathway [148]. The role of IL-17 as a pro-inflammatory cytokine in pancreatic cancer initiation and progression has not been investigated clearly and there are only a limited number of studies elucidating the factors contributing to it. IL-17, produced by the Th17 cells, have been found to be actively involved in causing the chronic inflammation leading to tumorigenesis [86, 167]. Human pancreatic cancer lesions have been found to be infiltrated by Th17 cells with overexpression of IL-17RA. Oncogenic KRAS and chronic pancreatitis populate Th17 cells and IL-17+ γδT cells at the tumor microenvironment. Mouse study with increase in expression of IL-17A indicated an increase in the occurrence of panIN [104]. REG-3β has been illustrated to promote KRAS mediated PDAC initiation, cell proliferation and inhibits

8

Role of Inflammatory Cytokines in the Initiation and Progression of. . .

145

apoptosis. REG-3β was found to increase the phosphorylation dependent activation of JAK2-STAT3 [30, 94]. IL-17 has been studied to increase pancreatic cancer stemness through the overexpression of couple of known cancer stemness genes (POU2F3 and DCLK1) along with ALDH1A1 [180]. Neutralizing antibodies against IL-17 and IL-17R effectively reduced the initiation and the progression of PanIN, indicating this cytokine as a therapeutic target for pancreatic cancer [104]. Monoclonal antibodies against IL-17 (Secukinumab and Ixekizumab) and IL-17R (Brodalumab-human monoclonal antibody) have already been employed in phase III trials to treat psoriasis. More potent antibodies to target IL-17 and better targets of IL-17 mediated pancreatic cancer could be developed with the better understanding of the underlying IL-17 signaling mechanism in this cancer [105].

8.6

Conclusion

Cytokines play an important role in the progression of pancreatic cancer. Several studies have reported the presence of different cytokines in the blood, tissue, pancreatic fluid. The PDAC patients have exhibited high amounts of IL-2, IL-6, IL-8 and IL-10 in serum. Cytokine signaling cascades are dysregulated in pancreatic cancer, causing inflammation leading to neoplasm. Both pro-inflammatory and antiinflammatory cytokines aid in the developing a tumor-promoting microenvironment helping the tumor to progress. So therapeutic strategies against these cytokines and their mechanisms are being investigated and developed. Current fundamental studies have showed positive results targeting cytokines like IL-6. Translating these findings in murine models to human studies are the next step to realize the goal of diagnosing the pancreatic cancer at an earlier stage.

References 1. Aldaz CM, Ferguson BW, Abba MC (2014) WWOX at the crossroads of cancer, metabolic syndrome related traits and CNS pathologies. Biochim Biophys Acta 1846(1):188–200. https://doi.org/10.1016/j.bbcan.2014.06.001 2. Anderson CF, Oukka M, Kuchroo VJ, Sacks D (2007) CD4(+)CD25(-)Foxp3(-) Th1 cells are the source of IL-10-mediated immune suppression in chronic cutaneous leishmaniasis. J Exp Med 204(2):285–297. https://doi.org/10.1084/jem.20061886 3. Ando K, Takahashi F, Motojima S, Nakashima K, Kaneko N, Hoshi K, Takahashi K (2013) Possible role for tocilizumab, an anti-interleukin-6 receptor antibody, in treating cancer cachexia. J Clin Oncol 31(6):e69–e72. https://doi.org/10.1200/JCO.2012.44.2020 4. Annes JP, Munger JS, Rifkin DB (2003) Making sense of latent TGFbeta activation. J Cell Sci 116(Pt 2):217–224 5. Anzano MA, Roberts AB, Meyers CA, Komoriya A, Lamb LC, Smith JM, Sporn MB (1982) Synergistic interaction of two classes of transforming growth factors from murine sarcoma cells. Cancer Res 42(11):4776–4778

146

M. A. Santharam and V. Dhandapani

6. Askling J, Fahrbach K, Nordstrom B, Ross S, Schmid CH, Symmons D (2011) Cancer risk with tumor necrosis factor alpha (TNF) inhibitors: meta-analysis of randomized controlled trials of adalimumab, etanercept, and infliximab using patient level data. Pharmacoepidemiol Drug Saf 20(2):119–130. https://doi.org/10.1002/pds.2046 7. Aykul S, Martinez-Hackert E (2016) Transforming Growth Factor-beta Family Ligands Can Function as Antagonists by Competing for Type II Receptor Binding. J Biol Chem 291 (20):10792–10804. https://doi.org/10.1074/jbc.M115.713487 8. Baeyens KJ, De Bondt HL, Raeymaekers A, Fiers W, De Ranter CJ (1999) The structure of mouse tumour-necrosis factor at 1.4 A resolution: towards modulation of its selectivity and trimerization. Acta Crystallogr D Biol Crystallogr 55(Pt 4):772–778 9. Balkwill F (2009) Tumour necrosis factor and cancer. Nat Rev Cancer 9(5):361–371. https:// doi.org/10.1038/nrc2628 10. Bao B, Ali S, Ahmad A, Azmi AS, Li Y, Banerjee S et al (2012) Hypoxia-induced aggressiveness of pancreatic cancer cells is due to increased expression of VEGF, IL-6 and miR-21, which can be attenuated by CDF treatment. PLoS One 7(12):e50165. https://doi.org/ 10.1371/journal.pone.0050165 11. Barry SP, Davidson SM, Townsend PA (2008) Molecular regulation of cardiac hypertrophy. Int J Biochem Cell Biol 40(10):2023–2039. https://doi.org/10.1016/j.biocel.2008.02.020 12. Batchu RB, Gruzdyn OV, Mahmud EM, Chukr F, Dachepalli R, Manmari SK et al (2018) Inhibition of Interleukin-10 in the tumor microenvironment can restore mesothelin chimeric antigen receptor T cell activity in pancreatic cancer in vitro. Surgery 163(3):627–632. https:// doi.org/10.1016/j.surg.2017.10.056 13. Bellone G, Carbone A, Smirne C, Scirelli T, Buffolino A, Novarino A et al (2006) Cooperative induction of a tolerogenic dendritic cell phenotype by cytokines secreted by pancreatic carcinoma cells. J Immunol 177(5):3448–3460 14. Bellone G, Smirne C, Mauri FA, Tonel E, Carbone A, Buffolino A et al (2006) Cytokine expression profile in human pancreatic carcinoma cells and in surgical specimens: implications for survival. Cancer Immunol Immunother 55(6):684–698. https://doi.org/10.1007/s00262005-0047-0 15. Benzel J, Fendrich V (2018) Familial Pancreatic Cancer. Oncol Res Treat 41(10):611–618. https://doi.org/10.1159/000493473 16. Berlato C, Cassatella MA, Kinjyo I, Gatto L, Yoshimura A, Bazzoni F (2002) Involvement of suppressor of cytokine signaling-3 as a mediator of the inhibitory effects of IL-10 on lipopolysaccharide-induced macrophage activation. J Immunol 168(12):6404–6411 17. Berti FCB, Pereira APL, Cebinelli GCM, Trugilo KP, Brajao de Oliveira K (2017) The role of interleukin 10 in human papilloma virus infection and progression to cervical carcinoma. Cytokine Growth Factor Rev 34:1–13. https://doi.org/10.1016/j.cytogfr.2017.03.002 18. Black RA, Rauch CT, Kozlosky CJ, Peschon JJ, Slack JL, Wolfson MF et al (1997) A metalloproteinase disintegrin that releases tumour-necrosis factor-alpha from cells. Nature 385(6618):729–733. https://doi.org/10.1038/385729a0 19. Blogowski W, Deskur A, Budkowska M, Salata D, Madej-Michniewicz A, Dabkowski K et al (2014) Selected cytokines in patients with pancreatic cancer: a preliminary report. PLoS One 9 (5):e97613. https://doi.org/10.1371/journal.pone.0097613 20. Bulek K, Liu C, Swaidani S, Wang L, Page RC, Gulen MF et al (2011) The inducible kinase IKKi is required for IL-17-dependent signaling associated with neutrophilia and pulmonary inflammation. Nat Immunol 12(9):844–852. https://doi.org/10.1038/ni.2080 21. Bynigeri RR, Jakkampudi A, Jangala R, Subramanyam C, Sasikala M, Rao GV et al (2017) Pancreatic stellate cell: Pandora’s box for pancreatic disease biology. World J Gastroenterol 23 (3):382–405. https://doi.org/10.3748/wjg.v23.i3.382 22. Cassatella MA, Gasperini S, Bovolenta C, Calzetti F, Vollebregt M, Scapini P et al (1999) Interleukin-10 (IL-10) selectively enhances CIS3/SOCS3 mRNA expression in human neutrophils: evidence for an IL-10-induced pathway that is independent of STAT protein activation. Blood 94(8):2880–2889

8

Role of Inflammatory Cytokines in the Initiation and Progression of. . .

147

23. Castro-Villegas C, Perez-Sanchez C, Escudero A, Filipescu I, Verdu M, Ruiz-Limon P et al (2015) Circulating miRNAs as potential biomarkers of therapy effectiveness in rheumatoid arthritis patients treated with anti-TNFalpha. Arthritis Res Ther 17:49. https://doi.org/10.1186/ s13075-015-0555-z 24. Chen G, Goeddel DV (2002) TNF-R1 signaling: a beautiful pathway. Science 296 (5573):1634–1635. https://doi.org/10.1126/science.1071924 25. Chen X, Tian J, Su GH, Lin J (2018) Blocking IL-6/GP130 signaling inhibits cell viability/ proliferation, glycolysis, and colony forming activity in human pancreatic cancer cells. Curr Cancer Drug Targets. https://doi.org/10.2174/1568009618666180430123939 26. Chen Y, Ayaru L, Mathew S, Morris E, Pereira SP, Behboudi S (2014) Expansion of antimesothelin specific CD4+ and CD8+ T cell responses in patients with pancreatic carcinoma. PLoS One 9(2):e88133. https://doi.org/10.1371/journal.pone.0088133 27. Cheng H, Fan K, Luo G, Fan Z, Yang C, Huang Q et al (2019) Kras(G12D) mutation contributes to regulatory T cell conversion through activation of the MEK/ERK pathway in pancreatic cancer. Cancer Lett 446:103–111. https://doi.org/10.1016/j.canlet.2019.01.013 28. Chow JY, Ban M, Wu HL, Nguyen F, Huang M, Chung H et al (2010) TGF-beta downregulates PTEN via activation of NF-kappaB in pancreatic cancer cells. Am J Physiol Gastrointest Liver Physiol 298(2):G275–G282. https://doi.org/10.1152/ajpgi.00344.2009 29. Ciechomska M, Bonek K, Merdas M, Zarecki P, Swierkot J, Gluszko P et al (2018) Changes in MiRNA-5196 Expression as a Potential Biomarker of Anti-TNF-alpha Therapy in Rheumatoid Arthritis and Ankylosing Spondylitis Patients. Arch Immunol Ther Exp (Warsz) 66 (5):389–397. https://doi.org/10.1007/s00005-018-0513-y 30. Corcoran RB, Contino G, Deshpande V, Tzatsos A, Conrad C, Benes CH et al (2011) STAT3 plays a critical role in KRAS-induced pancreatic tumorigenesis. Cancer Res 71 (14):5020–5029. https://doi.org/10.1158/0008-5472.CAN-11-0908 31. Couper KN, Blount DG, Riley EM (2008) IL-10: the master regulator of immunity to infection. J Immunol 180(9):5771–5777 32. Daley D, Zambirinis CP, Seifert L, Akkad N, Mohan N, Werba G et al (2016) γδ T cells support pancreatic oncogenesis by restraining αβ T cell activation. Cell 166(6):1485–1499. e1415 33. Demangel C, Bertolino P, Britton WJ (2002) Autocrine IL-10 impairs dendritic cell (DC)derived immune responses to mycobacterial infection by suppressing DC trafficking to draining lymph nodes and local IL-12 production. Eur J Immunol 32(4):994–1002. https:// doi.org/10.1002/1521-4141(200204)32:43.0.CO;2-6 34. DeSelm CJ, Tano ZE, Varghese AM, Adusumilli PS (2017) CAR T-cell therapy for pancreatic cancer. J Surg Oncol 116(1):63–74. https://doi.org/10.1002/jso.24627 35. Diakos CI, Charles KA, McMillan DC, Clarke SJ (2014) Cancer-related inflammation and treatment effectiveness. Lancet Oncol 15(11):e493–e503. https://doi.org/10.1016/S1470-2045 (14)70263-3 36. Dickens LS, Powley IR, Hughes MA, MacFarlane M (2012) The ‘complexities’ of life and death: death receptor signalling platforms. Exp Cell Res 318(11):1269–1277. https://doi.org/ 10.1016/j.yexcr.2012.04.005 37. Dijkgraaf EM, Heusinkveld M, Tummers B, Vogelpoel LT, Goedemans R, Jha V et al (2013) Chemotherapy alters monocyte differentiation to favor generation of cancer-supporting M2 macrophages in the tumor microenvironment. Cancer Res 73(8):2480–2492. https://doi.org/ 10.1158/0008-5472.CAN-12-3542 38. Ding WX, Yin XM (2004) Dissection of the multiple mechanisms of TNF-alpha-induced apoptosis in liver injury. J Cell Mol Med 8(4):445–454. https://doi.org/10.1111/j.1582-4934. 2004.tb00469.x 39. Ding Y, Chen D, Tarcsafalvi A, Su R, Qin L, Bromberg JS (2003) Suppressor of cytokine signaling 1 inhibits IL-10-mediated immune responses. J Immunol 170(3):1383–1391 40. Dranoff G (2004) Cytokines in cancer pathogenesis and cancer therapy. Nat Rev Cancer 4 (1):11–22. https://doi.org/10.1038/nrc1252

148

M. A. Santharam and V. Dhandapani

41. Dunn GP, Bruce AT, Ikeda H, Old LJ, Schreiber RD (2002) Cancer immunoediting: from immunosurveillance to tumor escape. Nat Immunol 3(11):991–998. https://doi.org/10.1038/ ni1102-991 42. Ebihara N, Matsuda A, Nakamura S, Matsuda H, Murakami A (2011) Role of the IL-6 classicand trans-signaling pathways in corneal sterile inflammation and wound healing. Invest Ophthalmol Vis Sci 52(12):8549–8557. https://doi.org/10.1167/iovs.11-7956 43. Ebrahimi B, Tucker SL, Li D, Abbruzzese JL, Kurzrock R (2004) Cytokines in pancreatic carcinoma: correlation with phenotypic characteristics and prognosis. Cancer 101 (12):2727–2736. https://doi.org/10.1002/cncr.20672 44. Egberts JH, Cloosters V, Noack A, Schniewind B, Thon L, Klose S et al (2008) Anti-tumor necrosis factor therapy inhibits pancreatic tumor growth and metastasis. Cancer Res 68 (5):1443–1450. https://doi.org/10.1158/0008-5472.CAN-07-5704 45. Ellenrieder V, Buck A, Harth A, Jungert K, Buchholz M, Adler G et al (2004) KLF11 mediates a critical mechanism in TGF-beta signaling that is inactivated by Erk-MAPK in pancreatic cancer cells. Gastroenterology 127(2):607–620 46. Ellenrieder V, Hendler SF, Ruhland C, Boeck W, Adler G, Gress TM (2001) TGF-betainduced invasiveness of pancreatic cancer cells is mediated by matrix metalloproteinase-2 and the urokinase plasminogen activator system. Int J Cancer 93(2):204–211. https://doi.org/10. 1002/ijc.1330 47. Feng L, Qi Q, Wang P, Chen H, Chen Z, Meng Z, Liu L (2018) Serum levels of IL-6, IL-8, and IL-10 are indicators of prognosis in pancreatic cancer. J Int Med Res 46(12):5228–5236. https://doi.org/10.1177/0300060518800588 48. Feng M, Xiong G, Cao Z, Yang G, Zheng S, Song X et al (2017) PD-1/PD-L1 and immunotherapy for pancreatic cancer. Cancer Lett 407:57–65. https://doi.org/10.1016/j. canlet.2017.08.006 49. Feurino LW, Fisher WE, Bharadwaj U, Yao Q, Chen C, Li M (2006) Current update of cytokines in pancreatic cancer: pathogenic mechanisms, clinical indication, and therapeutic values. Cancer Invest 24(7):696–703. https://doi.org/10.1080/07357900600981398 50. Feurino LW, Zhang Y, Bharadwaj U, Zhang R, Li F, Fisher WE et al (2007) IL-6 stimulates Th2 type cytokine secretion and upregulates VEGF and NRP-1 expression in pancreatic cancer cells. Cancer Biol Ther 6(7):1096–1100 51. Fiorentino DF, Bond MW, Mosmann TR (1989) Two types of mouse T helper cell. IV. Th2 clones secrete a factor that inhibits cytokine production by Th1 clones. J Exp Med 170 (6):2081–2095. https://doi.org/10.1084/jem.170.6.2081 52. Fiorentino DF, Zlotnik A, Mosmann TR, Howard M, O’Garra A (1991) IL-10 inhibits cytokine production by activated macrophages. J Immunol 147(11):3815–3822 53. Foulds KE, Rotte MJ, Seder RA (2006) IL-10 is required for optimal CD8 T cell memory following Listeria monocytogenes infection. J Immunol 177(4):2565–2574 54. Fujioka S, Nakamichi I, Esaki M, Asano K, Matsumoto T, Kitazono T (2014) Serum microRNA levels in patients with Crohn’s disease during induction therapy by infliximab. J Gastroenterol Hepatol 29(6):1207–1214. https://doi.org/10.1111/jgh.12523 55. Fukuda A, Wang SC, Morris JPT, Folias AE, Liou A, Kim GE et al (2011) Stat3 and MMP7 contribute to pancreatic ductal adenocarcinoma initiation and progression. Cancer Cell 19 (4):441–455. https://doi.org/10.1016/j.ccr.2011.03.002 56. Funamizu N, Hu C, Lacy C, Schetter A, Zhang G, He P et al (2013) Macrophage migration inhibitory factor induces epithelial to mesenchymal transition, enhances tumor aggressiveness and predicts clinical outcome in resected pancreatic ductal adenocarcinoma. Int J Cancer 132 (4):785–794. https://doi.org/10.1002/ijc.27736 57. Gahring LC, Carlson NG, Kulmar RA, Rogers SW (1996) Neuronal expression of tumor necrosis factor alpha in the murine brain. Neuroimmunomodulation 3(5):289–303. https://doi. org/10.1159/000097283

8

Role of Inflammatory Cytokines in the Initiation and Progression of. . .

149

58. Gao X, Cao Y, Yang W, Duan C, Aronson JF, Rastellini C et al (2013) BMP2 inhibits TGFbeta-induced pancreatic stellate cell activation and extracellular matrix formation. Am J Physiol Gastrointest Liver Physiol 304(9):G804–G813. https://doi.org/10.1152/ajpgi.00306. 2012 59. Gaur U, Aggarwal BB (2003) Regulation of proliferation, survival and apoptosis by members of the TNF superfamily. Biochem Pharmacol 66(8):1403–1408 60. Gilmore TD, Wolenski FS (2012) NF-kappaB: where did it come from and why? Immunol Rev 246(1):14–35. https://doi.org/10.1111/j.1600-065X.2012.01096.x 61. Giovarelli M, Musiani P, Modesti A, Dellabona P, Casorati G, Allione A et al (1995) Local release of IL-10 by transfected mouse mammary adenocarcinoma cells does not suppress but enhances antitumor reaction and elicits a strong cytotoxic lymphocyte and antibody-dependent immune memory. J Immunol 155(6):3112–3123 62. Gittler JK, Shemer A, Suarez-Farinas M, Fuentes-Duculan J, Gulewicz KJ, Wang CQ et al (2012) Progressive activation of T(H)2/T(H)22 cytokines and selective epidermal proteins characterizes acute and chronic atopic dermatitis. J Allergy Clin Immunol 130(6):1344–1354. https://doi.org/10.1016/j.jaci.2012.07.012 63. Harazono Y, Muramatsu T, Endo H, Uzawa N, Kawano T, Harada K et al (2013) miR-655 Is an EMT-suppressive microRNA targeting ZEB1 and TGFBR2. PLoS One 8(5):e62757. https://doi.org/10.1371/journal.pone.0062757 64. Hawa Z, Haque I, Ghosh A, Banerjee S, Harris L, Banerjee SK (2016) The miRacle in pancreatic cancer by miRNAs: tiny angels or devils in disease progression. Int J Mol Sci 17 (6). https://doi.org/10.3390/ijms17060809 65. Hayden MS, Ghosh S (2014) Regulation of NF-kappaB by TNF family cytokines. Semin Immunol 26(3):253–266. https://doi.org/10.1016/j.smim.2014.05.004 66. Heinrich PC, Behrmann I, Muller-Newen G, Schaper F, Graeve L (1998) Interleukin-6-type cytokine signalling through the gp130/Jak/STAT pathway. Biochem J 334(Pt 2):297–314. https://doi.org/10.1042/bj3340297 67. Heldin CH, Vanlandewijck M, Moustakas A (2012) Regulation of EMT by TGFbeta in cancer. FEBS Lett 586(14):1959–1970. https://doi.org/10.1016/j.febslet.2012.02.037 68. Herjan T, Yao P, Qian W, Li X, Liu C, Bulek K et al (2013) HuR is required for IL-17-induced Act1-mediated CXCL1 and CXCL5 mRNA stabilization. J Immunol 191(2):640–649. https:// doi.org/10.4049/jimmunol.1203315 69. Herreros-Villanueva M, Hijona E, Cosme A, Bujanda L (2012) Mouse models of pancreatic cancer. World J Gastroenterol 18(12):1286–1294. https://doi.org/10.3748/wjg.v18.i12.1286 70. Hofmann MA, Kiecker F, Zuberbier T (2016) A systematic review of the role of interleukin-17 and the interleukin-20 family in inflammatory allergic skin diseases. Curr Opin Allergy Clin Immunol 16(5):451–457. https://doi.org/10.1097/ACI.0000000000000310 71. Holmer R, Goumas FA, Waetzig GH, Rose-John S, Kalthoff H (2014) Interleukin-6: a villain in the drama of pancreatic cancer development and progression. Hepatobiliary Pancreat Dis Int 13(4):371–380 72. Huang C, Yang G, Jiang T, Huang K, Cao J, Qiu Z (2010) Effects of IL-6 and AG490 on regulation of Stat3 signaling pathway and invasion of human pancreatic cancer cells in vitro. J Exp Clin Cancer Res 29:51. https://doi.org/10.1186/1756-9966-29-51 73. Huang CK, Yang CY, Jeng YM, Chen CL, Wu HH, Chang YC et al (2014) Autocrine/ paracrine mechanism of interleukin-17B receptor promotes breast tumorigenesis through NFkappaB-mediated antiapoptotic pathway. Oncogene 33(23):2968–2977. https://doi.org/10. 1038/onc.2013.268 74. Hymowitz SG, Filvaroff EH, Yin JP, Lee J, Cai L, Risser P et al (2001) IL-17s adopt a cystine knot fold: structure and activity of a novel cytokine, IL-17F, and implications for receptor binding. EMBO J 20(19):5332–5341. https://doi.org/10.1093/emboj/20.19.5332 75. Ikushima H, Miyazono K (2010) TGFbeta signalling: a complex web in cancer progression. Nat Rev Cancer 10(6):415–424. https://doi.org/10.1038/nrc2853

150

M. A. Santharam and V. Dhandapani

76. Jang JE, Hajdu CH, Liot C, Miller G, Dustin ML, Bar-Sagi D (2017) Crosstalk between regulatory T cells and tumor-associated dendritic cells negates anti-tumor immunity in pancreatic cancer. Cell Rep 20(3):558–571. https://doi.org/10.1016/j.celrep.2017.06.062 77. Javle M, Li Y, Tan D, Dong X, Chang P, Kar S, Li D (2014) Biomarkers of TGF-beta signaling pathway and prognosis of pancreatic cancer. PLoS One 9(1):e85942. https://doi.org/ 10.1371/journal.pone.0085942 78. Johansson H, Andersson R, Bauden M, Hammes S, Holdenrieder S, Ansari D (2016) Immune checkpoint therapy for pancreatic cancer. World J Gastroenterol 22(43):9457–9476. https:// doi.org/10.3748/wjg.v22.i43.9457 79. Johnston FM, Tan MC, Tan BR Jr, Porembka MR, Brunt EM, Linehan DC et al (2009) Circulating mesothelin protein and cellular antimesothelin immunity in patients with pancreatic cancer. Clin Cancer Res 15(21):6511–6518. https://doi.org/10.1158/1078-0432.CCR-090565 80. Josefowicz SZ, Lu LF, Rudensky AY (2012) Regulatory T cells: mechanisms of differentiation and function. Annu Rev Immunol 30:531–564. https://doi.org/10.1146/annurev.immunol. 25.022106.141623 81. Joss A, Akdis M, Faith A, Blaser K, Akdis CA (2000) IL-10 directly acts on T cells by specifically altering the CD28 co-stimulation pathway. Eur J Immunol 30(6):1683–1690. https://doi.org/10.1002/1521-4141(200006)30:63.0.CO;2-A 82. Kabashima A, Higuchi H, Takaishi H, Matsuzaki Y, Suzuki S, Izumiya M et al (2009) Side population of pancreatic cancer cells predominates in TGF-beta-mediated epithelial to mesenchymal transition and invasion. Int J Cancer 124(12):2771–2779. https://doi.org/10.1002/ijc. 24349 83. Kang R, Loux T, Tang D, Schapiro NE, Vernon P, Livesey KM et al (2012) The expression of the receptor for advanced glycation endproducts (RAGE) is permissive for early pancreatic neoplasia. Proc Natl Acad Sci U S A 109(18):7031–7036. https://doi.org/10.1073/pnas. 1113865109 84. Kant S, Swat W, Zhang S, Zhang ZY, Neel BG, Flavell RA, Davis RJ (2011) TNF-stimulated MAP kinase activation mediated by a Rho family GTPase signaling pathway. Genes Dev 25 (19):2069–2078. https://doi.org/10.1101/gad.17224711 85. Karakhanova S, Link J, Heinrich M, Shevchenko I, Yang Y, Hassenpflug M et al (2015) Characterization of myeloid leukocytes and soluble mediators in pancreatic cancer: importance of myeloid-derived suppressor cells. Oncoimmunology 4(4):e998519. https://doi.org/10. 1080/2162402X.2014.998519 86. Kimura A, Naka T, Kishimoto T (2007) IL-6-dependent and -independent pathways in the development of interleukin 17-producing T helper cells. Proc Natl Acad Sci U S A 104 (29):12099–12104. https://doi.org/10.1073/pnas.0705268104 87. Klein AP, Brune KA, Petersen GM, Goggins M, Tersmette AC, Offerhaus GJ et al (2004) Prospective risk of pancreatic cancer in familial pancreatic cancer kindreds. Cancer Res 64 (7):2634–2638 88. Krintel SB, Dehlendorff C, Hetland ML, Horslev-Petersen K, Andersen KK, Junker P et al (2016) Prediction of treatment response to adalimumab: a double-blind placebo-controlled study of circulating microRNA in patients with early rheumatoid arthritis. Pharmacogenomics J 16(2):141–146. https://doi.org/10.1038/tpj.2015.30 89. Landskron G, De la Fuente M, Thuwajit P, Thuwajit C, Hermoso MA (2014) Chronic inflammation and cytokines in the tumor microenvironment. J Immunol Res 2014:149185. https://doi.org/10.1155/2014/149185 90. Lee J, Ho WH, Maruoka M, Corpuz RT, Baldwin DT, Foster JS et al (2001) IL-17E, a novel proinflammatory ligand for the IL-17 receptor homolog IL-17Rh1. J Biol Chem 276 (2):1660–1664. https://doi.org/10.1074/jbc.M008289200

8

Role of Inflammatory Cytokines in the Initiation and Progression of. . .

151

91. Lesina M, Kurkowski MU, Ludes K, Rose-John S, Treiber M, Kloppel G et al (2011) Stat3/ Socs3 activation by IL-6 transsignaling promotes progression of pancreatic intraepithelial neoplasia and development of pancreatic cancer. Cancer Cell 19(4):456–469. https://doi.org/ 10.1016/j.ccr.2011.03.009 92. Li MO, Wan YY, Sanjabi S, Robertson AK, Flavell RA (2006) Transforming growth factorbeta regulation of immune responses. Annu Rev Immunol 24:99–146. https://doi.org/10.1146/ annurev.immunol.24.021605.090737 93. Liang S, Ristich V, Arase H, Dausset J, Carosella ED, Horuzsko A (2008) Modulation of dendritic cell differentiation by HLA-G and ILT4 requires the IL-6--STAT3 signaling pathway. Proc Natl Acad Sci U S A 105(24):8357–8362. https://doi.org/10.1073/pnas. 0803341105 94. Loncle C, Bonjoch L, Folch-Puy E, Lopez-Millan MB, Lac S, Molejon MI et al (2015) IL17 functions through the novel REG3beta-JAK2-STAT3 inflammatory pathway to promote the transition from chronic pancreatitis to pancreatic cancer. Cancer Res 75(22):4852–4862. https://doi.org/10.1158/0008-5472.CAN-15-0896 95. Long KB, Tooker G, Tooker E, Luque SL, Lee JW, Pan X, Beatty GL (2017) IL6 Receptor Blockade Enhances Chemotherapy Efficacy in Pancreatic Ductal Adenocarcinoma. Mol Cancer Ther 16(9):1898–1908. https://doi.org/10.1158/1535-7163.MCT-16-0899 96. Lopez-Olivo MA, Tayar JH, Martinez-Lopez JA, Pollono EN, Cueto JP, Gonzales-Crespo MR et al (2012) Risk of malignancies in patients with rheumatoid arthritis treated with biologic therapy: a meta-analysis. JAMA 308(9):898–908. https://doi.org/10.1001/2012.jama.10857 97. Lowenfels AB, Maisonneuve P, Cavallini G, Ammann RW, Lankisch PG, Andersen JR et al (1993) Pancreatitis and the risk of pancreatic cancer. International Pancreatitis Study Group. N Engl J Med 328(20):1433–1437. https://doi.org/10.1056/NEJM199305203282001 98. Maezawa Y, Nakajima H, Suzuki K, Tamachi T, Ikeda K, Inoue J et al (2006) Involvement of TNF receptor-associated factor 6 in IL-25 receptor signaling. J Immunol 176(2):1013–1018 99. Mannino MH, Zhu Z, Xiao H, Bai Q, Wakefield MR, Fang Y (2015) The paradoxical role of IL-10 in immunity and cancer. Cancer Lett 367(2):103–107. https://doi.org/10.1016/j.canlet. 2015.07.009 100. Markowitz J, Brooks TR, Duggan MC, Paul BK, Pan X, Wei L et al (2015) Patients with pancreatic adenocarcinoma exhibit elevated levels of myeloid-derived suppressor cells upon progression of disease. Cancer Immunol Immunother 64(2):149–159. https://doi.org/10.1007/ s00262-014-1618-8 101. Masamune A, Yoshida N, Hamada S, Takikawa T, Nabeshima T, Shimosegawa T (2018) Exosomes derived from pancreatic cancer cells induce activation and profibrogenic activities in pancreatic stellate cells. Biochem Biophys Res Commun 495(1):71–77. https://doi.org/10. 1016/j.bbrc.2017.10.141 102. Massague J (2012) TGFbeta signalling in context. Nat Rev Mol Cell Biol 13(10):616–630. https://doi.org/10.1038/nrm3434 103. Masui T, Hosotani R, Doi R, Miyamoto Y, Tsuji S, Nakajima S et al (2002) Expression of IL-6 receptor in pancreatic cancer: involvement in VEGF induction. Anticancer Res 22 (6C):4093–4100 104. McAllister F, Bailey JM, Alsina J, Nirschl CJ, Sharma R, Fan H et al (2014) Oncogenic Kras activates a hematopoietic-to-epithelial IL-17 signaling axis in preinvasive pancreatic neoplasia. Cancer Cell 25(5):621–637. https://doi.org/10.1016/j.ccr.2014.03.014 105. McAllister F, Leach SD (2014) Targeting IL-17 for pancreatic cancer prevention. Oncotarget 5 (20):9530–9531. https://doi.org/10.18632/oncotarget.2618 106. McCarroll JA, Naim S, Sharbeen G, Russia N, Lee J, Kavallaris M et al (2014) Role of pancreatic stellate cells in chemoresistance in pancreatic cancer. Front Physiol 5:141. https:// doi.org/10.3389/fphys.2014.00141

152

M. A. Santharam and V. Dhandapani

107. Melisi D, Ishiyama S, Sclabas GM, Fleming JB, Xia Q, Tortora G et al (2008) LY2109761, a novel transforming growth factor beta receptor type I and type II dual inhibitor, as a therapeutic approach to suppressing pancreatic cancer metastasis. Mol Cancer Ther 7(4):829–840. https:// doi.org/10.1158/1535-7163.MCT-07-0337 108. Mitchell RE, Hassan M, Burton BR, Britton G, Hill EV, Verhagen J, Wraith DC (2017) IL-4 enhances IL-10 production in Th1 cells: implications for Th1 and Th2 regulation. Sci Rep 7 (1):11315. https://doi.org/10.1038/s41598-017-11803-y 109. Momi N, Kaur S, Krishn SR, Batra SK (2012) Discovering the route from inflammation to pancreatic cancer. Minerva Gastroenterol Dietol 58(4):283–297 110. Mondragon L, Kroemer G, Galluzzi L (2016) Immunosuppressive gammadelta T cells foster pancreatic carcinogenesis. Oncoimmunology 5(11):e1237328. https://doi.org/10.1080/ 2162402X.2016.1237328 111. Moore KW, de Waal Malefyt R, Coffman RL, O’Garra A (2001) Interleukin-10 and the interleukin-10 receptor. Annu Rev Immunol 19:683–765. https://doi.org/10.1146/annurev. immunol.19.1.683 112. Nakamura K, Kitani A, Strober W (2001) Cell contact-dependent immunosuppression by CD4 (+)CD25(+) regulatory T cells is mediated by cell surface-bound transforming growth factor beta. J Exp Med 194(5):629–644. https://doi.org/10.1084/jem.194.5.629 113. Ni CW, Hsieh HJ, Chao YJ, Wang DL (2004) Interleukin-6-induced JAK2/STAT3 signaling pathway in endothelial cells is suppressed by hemodynamic flow. Am J Physiol Cell Physiol 287(3):C771–C780. https://doi.org/10.1152/ajpcell.00532.2003 114. Nicolas FJ, Hill CS (2003) Attenuation of the TGF-beta-Smad signaling pathway in pancreatic tumor cells confers resistance to TGF-beta-induced growth arrest. Oncogene 22 (24):3698–3711. https://doi.org/10.1038/sj.onc.1206420 115. Novick D, Engelmann H, Wallach D, Rubinstein M (1989) Soluble cytokine receptors are present in normal human urine. J Exp Med 170(4):1409–1414. https://doi.org/10.1084/jem. 170.4.1409 116. O’Garra A, Vieira P (2004) Regulatory T cells and mechanisms of immune system control. Nat Med 10(8):801–805. https://doi.org/10.1038/nm0804-801 117. O’Shea JJ, Holland SM, Staudt LM (2013) JAKs and STATs in immunity, immunodeficiency, and cancer. N Engl J Med 368(2):161–170. https://doi.org/10.1056/NEJMra1202117 118. Oeckinghaus A, Hayden MS, Ghosh S (2011) Crosstalk in NF-kappaB signaling pathways. Nat Immunol 12(8):695–708. https://doi.org/10.1038/ni.2065 119. Oida T, Weiner HL (2010) Overexpression of TGF-ss 1 gene induces cell surface localized glucose-regulated protein 78-associated latency-associated peptide/TGF-ss. J Immunol 185 (6):3529–3535. https://doi.org/10.4049/jimmunol.0904121 120. Olszewski MB, Groot AJ, Dastych J, Knol EF (2007) TNF trafficking to human mast cell granules: mature chain-dependent endocytosis. J Immunol 178(9):5701–5709 121. Padua D, Massague J (2009) Roles of TGFbeta in metastasis. Cell Res 19(1):89–102. https:// doi.org/10.1038/cr.2008.316 122. Palladino MA, Bahjat FR, Theodorakis EA, Moldawer LL (2003) Anti-TNF-alpha therapies: the next generation. Nat Rev Drug Discov 2(9):736–746. https://doi.org/10.1038/nrd1175 123. Pinho AV, Chantrill L, Rooman I (2014) Chronic pancreatitis: a path to pancreatic cancer. Cancer Lett 345(2):203–209. https://doi.org/10.1016/j.canlet.2013.08.015 124. Pivarcsi A, Meisgen F, Xu N, Stahle M, Sonkoly E (2013) Changes in the level of serum microRNAs in patients with psoriasis after antitumour necrosis factor-alpha therapy. Br J Dermatol 169(3):563–570. https://doi.org/10.1111/bjd.12381 125. Qian Y, Liu C, Hartupee J, Altuntas CZ, Gulen MF, Jane-Wit D et al (2007) The adaptor Act1 is required for interleukin 17-dependent signaling associated with autoimmune and inflammatory disease. Nat Immunol 8(3):247–256. https://doi.org/10.1038/ni1439 126. Qu F, Gao H, Zhu S, Shi P, Zhang Y, Liu Y et al (2012) TRAF6-dependent Act1 phosphorylation by the IkappaB kinase-related kinases suppresses interleukin-17-induced NF-kappaB activation. Mol Cell Biol 32(19):3925–3937. https://doi.org/10.1128/MCB.00268-12

8

Role of Inflammatory Cytokines in the Initiation and Progression of. . .

153

127. Ramirez-Carrozzi V, Sambandam A, Luis E, Lin Z, Jeet S, Lesch J et al (2011) IL-17C regulates the innate immune function of epithelial cells in an autocrine manner. Nat Immunol 12(12):1159–1166. https://doi.org/10.1038/ni.2156 128. Razidlo GL, Burton KM, McNiven MA (2018) Interleukin-6 promotes pancreatic cancer cell migration by rapidly activating the small GTPase CDC42. J Biol Chem 293 (28):11143–11153. https://doi.org/10.1074/jbc.RA118.003276 129. Rickel EA, Siegel LA, Yoon BR, Rottman JB, Kugler DG, Swart DA et al (2008) Identification of functional roles for both IL-17RB and IL-17RA in mediating IL-25-induced activities. J Immunol 181(6):4299–4310 130. Rifkin DB (2005) Latent transforming growth factor-beta (TGF-beta) binding proteins: orchestrators of TGF-beta availability. J Biol Chem 280(9):7409–7412. https://doi.org/10. 1074/jbc.R400029200 131. Ripka S, Konig A, Buchholz M, Wagner M, Sipos B, Kloppel G et al (2007) WNT5A--target of CUTL1 and potent modulator of tumor cell migration and invasion in pancreatic cancer. Carcinogenesis 28(6):1178–1187. https://doi.org/10.1093/carcin/bgl255 132. Rose-John S (2012) IL-6 trans-signaling via the soluble IL-6 receptor: importance for the pro-inflammatory activities of IL-6. Int J Biol Sci 8(9):1237–1247. https://doi.org/10.7150/ ijbs.4989 133. Rose-John S, Heinrich PC (1994) Soluble receptors for cytokines and growth factors: generation and biological function. Biochem J 300(Pt 2):281–290. https://doi.org/10.1042/ bj3000281 134. Rose-John S, Waetzig GH, Scheller J, Grotzinger J, Seegert D (2007) The IL-6/sIL-6R complex as a novel target for therapeutic approaches. Expert Opin Ther Targets 11 (5):613–624. https://doi.org/10.1517/14728222.11.5.613 135. Roshani R, McCarthy F, Hagemann T (2014) Inflammatory cytokines in human pancreatic cancer. Cancer Lett 345(2):157–163. https://doi.org/10.1016/j.canlet.2013.07.014 136. Rouvier E, Luciani MF, Mattei MG, Denizot F, Golstein P (1993) CTLA-8, cloned from an activated T cell, bearing AU-rich messenger RNA instability sequences, and homologous to a herpesvirus saimiri gene. J Immunol 150(12):5445–5456 137. Saarma M (2000) GDNF – a stranger in the TGF-beta superfamily? Eur J Biochem 267 (24):6968–6971 138. Saraiva M, Christensen JR, Veldhoen M, Murphy TL, Murphy KM, O’Garra A (2009) Interleukin-10 production by Th1 cells requires interleukin-12-induced STAT4 transcription factor and ERK MAP kinase activation by high antigen dose. Immunity 31(2):209–219. https://doi.org/10.1016/j.immuni.2009.05.012 139. Schandene L, Alonso-Vega C, Willems F, Gerard C, Delvaux A, Velu T et al (1994) B7/ CD28-dependent IL-5 production by human resting T cells is inhibited by IL-10. J Immunol 152(9):4368–4374 140. Scheller J, Garbers C, Rose-John S (2014) Interleukin-6: from basic biology to selective blockade of pro-inflammatory activities. Semin Immunol 26(1):2–12. https://doi.org/10. 1016/j.smim.2013.11.002 141. Schust J, Sperl B, Hollis A, Mayer TU, Berg T (2006) Stattic: a small-molecule inhibitor of STAT3 activation and dimerization. Chem Biol 13(11):1235–1242. https://doi.org/10.1016/j. chembiol.2006.09.018 142. Schwandner R, Yamaguchi K, Cao Z (2000) Requirement of tumor necrosis factor receptorassociated factor (TRAF)6 in interleukin 17 signal transduction. J Exp Med 191 (7):1233–1240. https://doi.org/10.1084/jem.191.7.1233 143. Shadhu K, Xi C (2019) Inflammation and pancreatic cancer: An updated review. Saudi J Gastroenterol 25(1):3–13. https://doi.org/10.4103/sjg.SJG_390_18 144. Shah N, Kammermeier J, Elawad M, Glocker EO (2012) Interleukin-10 and interleukin-10receptor defects in inflammatory bowel disease. Curr Allergy Asthma Rep 12(5):373–379. https://doi.org/10.1007/s11882-012-0286-z

154

M. A. Santharam and V. Dhandapani

145. Shek FW, Benyon RC, Walker FM, McCrudden PR, Pender SL, Williams EJ et al (2002) Expression of transforming growth factor-beta 1 by pancreatic stellate cells and its implications for matrix secretion and turnover in chronic pancreatitis. Am J Pathol 160 (5):1787–1798 146. Shinriki S, Jono H, Ota K, Ueda M, Kudo M, Ota T et al (2009) Humanized anti-interleukin-6 receptor antibody suppresses tumor angiogenesis and in vivo growth of human oral squamous cell carcinoma. Clin Cancer Res 15(17):5426–5434. https://doi.org/10.1158/1078-0432.CCR09-0287 147. Siegel RL, Miller KD, Jemal A (2018) Cancer statistics, 2018. CA Cancer J Clin 68(1):7–30. https://doi.org/10.3322/caac.21442 148. Song X, Zhu S, Shi P, Liu Y, Shi Y, Levin SD, Qian Y (2011) IL-17RE is the functional receptor for IL-17C and mediates mucosal immunity to infection with intestinal pathogens. Nat Immunol 12(12):1151–1158. https://doi.org/10.1038/ni.2155 149. Sun D, Novotny M, Bulek K, Liu C, Li X, Hamilton T (2011) Treatment with IL-17 prolongs the half-life of chemokine CXCL1 mRNA via the adaptor TRAF5 and the splicing-regulatory factor SF2 (ASF). Nat Immunol 12(9):853–860. https://doi.org/10.1038/ni.2081 150. Taga T, Kishimoto T (1997) Gp130 and the interleukin-6 family of cytokines. Annu Rev Immunol 15:797–819. https://doi.org/10.1146/annurev.immunol.15.1.797 151. Takano S, Kanai F, Jazag A, Ijichi H, Yao J, Ogawa H et al (2007) Smad4 is essential for down-regulation of E-cadherin induced by TGF-beta in pancreatic cancer cell line PANC-1. J Biochem 141(3):345–351. https://doi.org/10.1093/jb/mvm039 152. Tanaka T, Narazaki M, Kishimoto T (2014) IL-6 in inflammation, immunity, and disease. Cold Spring Harb Perspect Biol 6(10):a016295. https://doi.org/10.1101/cshperspect.a016295 153. Tang RF, Wang SX, Zhang FR, Peng L, Wang SX, Xiao Y, Zhang M (2005) Interleukin1alpha, 6 regulate the secretion of vascular endothelial growth factor A, C in pancreatic cancer. Hepatobiliary Pancreat Dis Int 4(3):460–463 154. Thannickal VJ, Fanburg BL (2000) Reactive oxygen species in cell signaling. Am J Physiol Lung Cell Mol Physiol 279(6):L1005–L1028. https://doi.org/10.1152/ajplung.2000.279.6. L1005 155. Theiss AL, Simmons JG, Jobin C, Lund PK (2005) Tumor necrosis factor (TNF) alpha increases collagen accumulation and proliferation in intestinal myofibroblasts via TNF receptor 2. J Biol Chem 280(43):36099–36109. https://doi.org/10.1074/jbc.M505291200 156. Trinchieri G (2007) Interleukin-10 production by effector T cells: Th1 cells show self control. J Exp Med 204(2):239–243. https://doi.org/10.1084/jem.20070104 157. Vasseur P, Devaure I, Sellier J, Delwail A, Chagneau-Derrode C, Charier F et al (2014) High plasma levels of the pro-inflammatory cytokine IL-22 and the anti-inflammatory cytokines IL-10 and IL-1ra in acute pancreatitis. Pancreatology 14(6):465–469. https://doi.org/10.1016/ j.pan.2014.08.005 158. Villanueva A, Garcia C, Paules AB, Vicente M, Megias M, Reyes G et al (1998) Disruption of the antiproliferative TGF-beta signaling pathways in human pancreatic cancer cells. Oncogene 17(15):1969–1978. https://doi.org/10.1038/sj.onc.1202118 159. von Ahrens D, Bhagat TD, Nagrath D, Maitra A, Verma A (2017) The role of stromal cancerassociated fibroblasts in pancreatic cancer. J Hematol Oncol 10(1):76. https://doi.org/10.1186/ s13045-017-0448-5 160. Wajant H, Pfizenmaier K, Scheurich P (2003) Tumor necrosis factor signaling. Cell Death Differ 10(1):45–65. https://doi.org/10.1038/sj.cdd.4401189 161. Waters JP, Pober JS, Bradley JR (2013) Tumour necrosis factor and cancer. J Pathol 230 (3):241–248. https://doi.org/10.1002/path.4188 162. Wey JS, Gray MJ, Fan F, Belcheva A, McCarty MF, Stoeltzing O et al (2005) Overexpression of neuropilin-1 promotes constitutive MAPK signalling and chemoresistance in pancreatic cancer cells. Br J Cancer 93(2):233–241. https://doi.org/10.1038/sj.bjc.6602663

8

Role of Inflammatory Cytokines in the Initiation and Progression of. . .

155

163. Williams CJ, Peyrin-Biroulet L, Ford AC (2014) Systematic review with meta-analysis: malignancies with anti-tumour necrosis factor-alpha therapy in inflammatory bowel disease. Aliment Pharmacol Ther 39(5):447–458. https://doi.org/10.1111/apt.12624 164. Williams L, Bradley L, Smith A, Foxwell B (2004) Signal transducer and activator of transcription 3 is the dominant mediator of the anti-inflammatory effects of IL-10 in human macrophages. J Immunol 172(1):567–576 165. Wipff PJ, Hinz B (2008) Integrins and the activation of latent transforming growth factor beta1 – an intimate relationship. Eur J Cell Biol 87(8-9):601–615. https://doi.org/10.1016/j.ejcb. 2008.01.012 166. Wu L, Zepp JA, Qian W, Martin BN, Ouyang W, Yin W et al (2015) A novel IL-25 signaling pathway through STAT5. J Immunol 194(9):4528–4534. https://doi.org/10.4049/jimmunol. 1402760 167. Wu S, Rhee KJ, Albesiano E, Rabizadeh S, Wu X, Yen HR et al (2009) A human colonic commensal promotes colon tumorigenesis via activation of T helper type 17 T cell responses. Nat Med 15(9):1016–1022. https://doi.org/10.1038/nm.2015 168. Xing HB, Tong MT, Wang J, Hu H, Zhai CY, Huang CX, Li D (2018) Suppression of IL-6 gene by shRNA augments gemcitabine chemosensitization in pancreatic adenocarcinoma cells. Biomed Res Int 2018:3195025. https://doi.org/10.1155/2018/3195025 169. Xu XF, Liu F, Xin JQ, Fan JW, Wu N, Zhu LJ et al (2018) Respective roles of the mitogenactivated protein kinase (MAPK) family members in pancreatic stellate cell activation induced by transforming growth factor-beta1 (TGF-beta1). Biochem Biophys Res Commun 501 (2):365–373. https://doi.org/10.1016/j.bbrc.2018.04.176 170. Yagil Z, Nechushtan H, Kay G, Yang CM, Kemeny DM, Razin E (2010) The enigma of the role of protein inhibitor of activated STAT3 (PIAS3) in the immune response. Trends Immunol 31(5):199–204. https://doi.org/10.1016/j.it.2010.01.005 171. Yako YY, Kruger D, Smith M, Brand M (2016) Cytokines as biomarkers of pancreatic ductal adenocarcinoma: a systematic review. PLoS One 11(5):e0154016. https://doi.org/10.1371/ journal.pone.0154016 172. Yang Z, Ren F, Liu C, He S, Sun G, Gao Q et al (2010) dbDEMC: a database of differentially expressed miRNAs in human cancers. BMC Genomics 11(Suppl 4):S5. https://doi.org/10. 1186/1471-2164-11-S4-S5 173. Yao X, Huang J, Zhong H, Shen N, Faggioni R, Fung M, Yao Y (2014) Targeting interleukin6 in inflammatory autoimmune diseases and cancers. Pharmacol Ther 141(2):125–139. https:// doi.org/10.1016/j.pharmthera.2013.09.004 174. Yao Z, Fanslow WC, Seldin MF, Rousseau AM, Painter SL, Comeau MR et al (1995) Herpesvirus Saimiri encodes a new cytokine, IL-17, which binds to a novel cytokine receptor. Immunity 3(6):811–821 175. Yao Z, Painter SL, Fanslow WC, Ulrich D, Macduff BM, Spriggs MK, Armitage RJ (1995) Human IL-17: a novel cytokine derived from T cells. J Immunol 155(12):5483–5486 176. Yu H, Pardoll D, Jove R (2009) STATs in cancer inflammation and immunity: a leading role for STAT3. Nat Rev Cancer 9(11):798–809. https://doi.org/10.1038/nrc2734 177. Zepp JA, Wu L, Qian W, Ouyang W, Aronica M, Erzurum S, Li X (2015) TRAF4-SMURF2mediated DAZAP2 degradation is critical for IL-25 signaling and allergic airway inflammation. J Immunol 194(6):2826–2837. https://doi.org/10.4049/jimmunol.1402647 178. Zhang Y, Bharadwaj U, Logsdon CD, Chen C, Yao Q, Li M (2010) ZIP4 regulates pancreatic cancer cell growth by activating IL-6/STAT3 pathway through zinc finger transcription factor CREB. Clin Cancer Res 16(5):1423–1430. https://doi.org/10.1158/1078-0432.CCR-09-2405 179. Zhang Y, Yan W, Collins MA, Bednar F, Rakshit S, Zetter BR et al (2013) Interleukin-6 is required for pancreatic cancer progression by promoting MAPK signaling activation and oxidative stress resistance. Cancer Res 73(20):6359–6374. https://doi.org/10.1158/00085472.CAN-13-1558-T

156

M. A. Santharam and V. Dhandapani

180. Zhang Y, Zoltan M, Riquelme E, Xu H, Sahin I, Castro-Pando S et al (2018) Immune cell production of interleukin 17 induces stem cell features of pancreatic intraepithelial neoplasia cells. Gastroenterology 155(1):210–223. e213. https://doi.org/10.1053/j.gastro.2018.03.041 181. Zhao X, Fan W, Xu Z, Chen H, He Y, Yang G et al (2016) Inhibiting tumor necrosis factoralpha diminishes desmoplasia and inflammation to overcome chemoresistance in pancreatic ductal adenocarcinoma. Oncotarget 7(49):81110–81122. https://doi.org/10.18632/oncotarget. 13212 182. Zrioual S, Ecochard R, Tournadre A, Lenief V, Cazalis MA, Miossec P (2009) Genome-wide comparison between IL-17A- and IL-17F-induced effects in human rheumatoid arthritis synoviocytes. J Immunol 182(5):3112–3120. https://doi.org/10.4049/jimmunol.0801967

9

Perspectives and Molecular Understanding of Pancreatic Cancer Stem Cells L. Saikrishna, Prameswari Kasa, Saimila Momin, and L. V. K. S. Bhaskar

Abstract

Pancreatic cancer (PC) is a lethal, malignant cancer that bears high mortality rates. Due to its lack of noticeable symptoms, it is often diagnosed late and current pancreatic cancer therapies are ineffective with poor prognosis. Despite the high mortality and poor survival of PC patients, there is limited information on factors propagating resistance. Resistance to standard therapies in pancreatic cancer patients is partly associated with the presence of a subpopulation of highly plastic “stem-like” cells (pancreatic cancer stem cells: paCSC) in tumors. In this connection, it is important to have a strong understanding of the paCSC population, especially its distinct characteristics, in order to engineer new therapies to target these resistant cells. Therefore, the purpose of this investigation is to highlight and discuss PaCSC and their specific surface markers. Overall, in this study, we searched MEDLINE, EMBASE, the Cochrane Library, Web of Science, and ISI Proceedings for observational studies relating to the PaCSC and PC. Pancreatic cancer stem cells exhibit specific immune characteristics on their surface. The CD133, CD44, CD24, ALDH1, c-Met, DCLK1, CXCR4, EpCAM and ABCG2 are prominent Pa-CSC markers. As PaCSCs drive tumorigenesis and metastasis, their manipulation approaches would have widespread clinical implications and hence improve outcomes in pancreatic cancer. L. Saikrishna Department of Zoology, Visvodaya Government Degree College, Venkatagiri, India P. Kasa Dr. LV Prasad Diagnostics and Research Laboratory, Hyderabad, Telangana, India S. Momin Department of Hematology and Medical Oncology, Winship Cancer Institute, Emory University, Atlanta, GA, USA L. V. K. S. Bhaskar (*) Guru Ghasidas Vishwavidyalaya, Bilaspur, India e-mail: [email protected] # Springer Nature Singapore Pte Ltd. 2019 G. P. Nagaraju, A. BM Reddy (eds.), Exploring Pancreatic Metabolism and Malignancy, https://doi.org/10.1007/978-981-32-9393-9_9

157

158

L. Saikrishna et al.

Keywords

Cancer stem cells · Pancreatic cancer · Stem cell markers · Signaling cascades

Abbreviations 15-PGDH 5-FU ABCG2 ADLH ALCAM CSCs CXCR4 DCLK1 EMT EpCAM HA HGF mRNA nab-paclitaxel NETs NOD/SCID PaCSCs PC PDAC PDX1 SP TME

9.1

15-Hydroxyprostaglandin dehydrogenase 5-fluorouracil ATP-binding cassette subfamily G member 2 Aldehyde dehydrogenase leukocyte cell adhesion molecule Cancer stem cells C-X-C chemokine receptor type 4 Doublecortin-like kinase 1 Epithelial mesenchymal transition Epithelial cell adhesion molecule hyaluronic acid Hepatocyte growth factor Messenger RNA albumin-bound paclitaxel Neuroendocrine tumors nonobese diabetic/severe combined immunodeficiency Pancreatic cancer stem cells Pancreatic cancer pancreatic ductal adenocarcinoma pancreatic and duodenal homeobox 1 side-population Tumor microenvironment

Introduction

Pancreatic cancer (PC) is considered to be one of the most lethal cancers and is currently the seventh leading cause of global cancer deaths [69]. It is the most aggressive and hazardous malignancy with high rates of mortality [1]. Over the past years, the incidence rates have increased for pancreatic cancer. The individuals with risk factors such as tobacco usage, obesity, diabetes and pancreatitis, and positive family history were more susceptible to the risk of PC [55]. About 5–10% of cases with PC are genetic with about 80% penetrance [4, 14]. Although the exact reason for the differences in the incidence of PC among global populations is not known, the variations in certain risk factors such as tobacco smoking, dietary style and obesity could account for these differences [26, 38]. Furthermore, the incidence and mortality of PC in global populations also correlated with increasing age

9

Perspectives and Molecular Understanding of Pancreatic Cancer Stem Cells

159

[12]. Considerable variations (30-fold) in the mortality rates for PC in the global populations include: highest mortality in European countries and lowest in Eastern African and South-Eastern Asian countries [12]. Survival rates of PC are determined by the stage at the time of diagnosis, tumor size, healthcare facilities available and completeness of follow-up [92]. However, over the past 35 years, the survival rate of PC has increased from 3% to 6.8% [5]. Standard treatments currently for PC include 5-fluorouracil (5-FU), gemcitabine, FOLFIRINOX and nab-paclitaxel (albuminbound paclitaxel). However, these treatments only extend the survival of patients up to a year and often fail to eradicate tumors. Additionally, pancreatic cancer is highly resistant to both radiotherapy and chemotherapy and has the ability to quickly spread to other organs, causing irreparable damage. Current studies and advancements have definitely contributed to our understanding of the pathophysiology of pancreatic cancer, specifically on the causes of cancer metastasis, PC relapse, and resistance towards chemotherapy Along these lines, a subset of cancer cells with stem-like properties that coincide with other cellular units/factors of tumor microenvironment (TME) have been identified. These cells are described as cancer stem cells (CSCs), the key drivers of tumor initiation and progression.

9.2

Cancer Stem Cells

Stem cells are described as subpopulations of cancer cells with distinct properties such as self-renewal and multi-lineage differentiation. Formation of CSC and tumor development are driven mainly by the reactivation of certain embryonic signaling pathways in cancer cells [44, 95]. Although, epigenetic and metabolic alterations in cancer cells are highly intertwined, CSCs are both epigenetically and metabolically different from non-CSCs of the surrounding niche environment. Hence effective recognition and characterization of CSCs is possible through the expression of specific cell surface markers. The existence of specific CSC clones depends on several extrinsic factors such as hypoxia, chemotherapy, and even nutrient deprivation [86]. Hence, the heterogeneous cell populations derived from these CSCs determine tumor heterogeneity [66]. Further mutations in stem cells were found to increase cancer risk by causing interruptions in cell division or preventing differentiation, which led to a significant increase in stem cell proliferation [99]. Understanding of CSCs may improve existing therapies and help us in finding new therapeutic methods against CSCs. It would seem that the new manipulation approaches of cancer stem cells would have widespread clinical implications and hence improve outcomes in pancreatic cancer.

9.3

Markers of PC Stem Cells

The presence of CSCs in various solid tumors was demonstrated after two decades of the identification of CSCs in human acute myeloid leukemia [11, 54, 80]. Subsequent studies identified various cell surface markers expressed by pancreatic cancer stem

160

L. Saikrishna et al.

Fig. 9.1 Prominent pancreatic cancer stem cell markers

cells (PaCSCs). Some markers are of specific functional relevance to PC (Fig. 9.1). The prominent Pa-CSC markers include CD133, CD44, CD24, ALDH1, c-Met, DCLK1, CXCR4, EpCAM and ABCG2 (Table 9.1).

9.3.1

CD133

CD133 (prominin-1) is a glycoprotein that serves as a CSC marker in a large variety of human malignancies [28, 41]. The CD133 is encoded by PROM1 gene on human chromosome 4. Transcription of this gene is carefully regulated in a tissue-specific manner; in other words, expression is only found in certain tissues [21]. According to studies on CD133 mRNAs, these tissues include pancreas, testis, kidney, placenta, mammary gland and digestive tract [23]. Additionally, CD133 is associated with the Notch pathway and Hedgehog signaling, such that when these pathways are inhibited (i.e. Sonic hedgehog, Shh, is inhibited), there is an increased sensitivity of CD133+ glioma cells towards temozolomide therapy [84]. CD133 plays a strong role in motility, invasiveness and epithelial mesenchymal transition (EMT) due to their location in internalization prone membrane domains and involvement in multiple signaling pathways. Hypoxia induces the expansion of CD133+ that leads to tumor aggressiveness of pancreatic cancer cells in the presence of hypoxiainducible factor 1-alpha (HIF-1α) [29]. Pancreatic cancer tissue was shown to have CD133+, highly tumorigenic and gemcitabine resistant migrating cells. Depletion of CD133+ and CXCR4+ pancreatic cancer cells significantly decreased the tumorigenic potential of PC [30]. CD133 is localized to areas rich in cadherin and associated with the signaling pathways regulating the EMT program. Downregulation of endogenous CD133 in the Capan1M9 cells contributed towards Slug

9

Perspectives and Molecular Understanding of Pancreatic Cancer Stem Cells

161

Table 9.1 Markers of pancreatic cancer stem cells Cell surface marker CD44

Characteristics Non-kinase transmembrane glycoprotein

Function Cell adhesion molecule, receptor for hyaluronic acid

CD24

Glycosyl phosphatidylinositol (GPI)-linked sialoglycoprotein

Cell adhesion molecule

CD133

Penta-span transmembrane glycoprotein.

Organizer of cell membrane topology.

ALDH1

Nicotinamideadenine dinucleotide phosphate–positive (NAD(P)+)dependent enzymes

Cellular detoxifying enzyme

c-met

Tyrosine-protein kinase met

Hepatocyte growth factor receptor (HGFR)

DCLK1

Microtubuleassociated kinase

Involved in the epithelialmesenchymal transition (EMT).

Significance Cell adhesion, migration, drug resistance and apoptosis. Knockdown of CD44 expression by its shRNA in pancreatic cancer cells led to decreased cellular proliferation and migration ability. Enhanced clonogenic capacity, multilineage potential, invasiveness, high proliferation and poor prognosis. Self-renewal pathway signaling, increased tumorigenicity, tumor progression, and metastasis. Display stem-like features, such as selfrenewal, clonogenic growth, tumorinitiating capacity. Knockdown of ALDH1B1 caused a 35% reduction in cell growth in the high ALDH1B1expressing cell lines. Signaling induces growth and invasion. Cells with increased c-met has tumorigenic potential. DCLK1 is essential for the invasive and metastatic properties of CSC. Overexpression promotes the tumor progression, sphere formation and migration of PC cells. Knockdown of

Reference [49, 97]

[53, 62, 71]

[28, 34, 61]

[75, 82, 87]

[31, 47, 93]

[36, 50]

(continued)

162

L. Saikrishna et al.

Table 9.1 (continued) Cell surface marker

Characteristics

Function

CXCR4

Seven transmembrane Gprotein-linked CXC chemokine receptor

It is an alphachemokine receptor specific for stromalderived-factor-1 (SDF-1 or CXCL12)

EpCAM

Transmembrane glycoprotein

Mediate Ca2+ independent homotypic cell-cell adhesion in epithelia.

ABCG2

ATP-binding cassette (ABC) transporter

Drug efflux pump

Significance DCLK1 suppressed liver metastasis of pancreatic CSCs. Involved in the occurrence of metastasis in PC. CXCR4/ CXCL12 axis involved in cell proliferation, migration, and angiogenesis. Cell signaling, cellcell adhesion, proliferation, tumorigenesis, differentiation, metastasis and migration. ABCG2 overexpression lead to fast progression of malignancy, unsuccessful chemotherapy and poor prognosis.

Reference

[13, 30, 89]

[52, 64]

[88, 94]

suppression and a decrease in migration/motility and invasion [19]. In patients who have tumors that are not excisable, the ascites-derived exosomes of these patients suggested that highly glycosylated CD133 is a possible biomarker for improved prognosis of patients [70]. Furthermore, an overlap of 10–40% between CD133+ and CD44+, CD24+, EpCAM+ was reported in PDAC cells [30].

9.3.2

CD44

CD44 is a cell surface glycoprotein that mediates interactions between neighboring cells, cellular adhesion and cellular migration. CD44 gene produces a variety of isoforms (CD44s and CD44v) that are expressed differently in both normal and cancer cells. Analysis of CD44 isoforms in primary and metastatic human pancreatic adenocarcinomas revealed a novel CD44(v6) isoform in metastatic lesions [67]. Furthermore, CD44v6 expression was more prevalent in human tumor tissues of advanced stages of metastatic pancreatic carcinomas [48]. Pancreatic cancer cells (MiaPaCa-2) are involved in hyaluronic acid (HA) and HA oligosaccharide

9

Perspectives and Molecular Understanding of Pancreatic Cancer Stem Cells

163

generation. Tumor-derived HA oligosaccharides can increase CD44 intracellular cleavage and motility of tumor cells; when this interaction is inhibited it can lead to a halt in tumor cell motility [77]. Knockdown of CD44 expression reduces cellular expansion and migration in PC cells when there is a simultaneous reduction in p-ERK and p-AKT concentrations [49]. Overexpression of CD44 in pancreatic cancer cell lines and pancreatic tumors indicating the CD44 plays a major role in carcinogenesis and progression of pancreatic cancer [43]. Additionally, correlation of clinicopathological factors with CD44 expression demonstrated that high CD44 expression predicts early recurrence of tumors in PC patients undergoing radical surgery [42]. The ionophore antibiotic gramicidin reduced the expression levels of CD133, CD44, and CD47 and suppressed cancer cell proliferation in synergism with gemcitabine [90]. In addition, BRM270 also inhibits metastasis and stem-like traits in CD44+ pancreatic ductal adenocarcinoma (PDAC) through the Shh signaling pathway [33].

9.3.3

CD24

CD24 is a heat stable antigen and a biomarker in both healthy and malignant stem cells [38, 81]. CD24 is a surface marker for pancreatic and duodenal homeobox 1 (PDX1) positive pancreatic progenitors derived from human embryonic stem cells [39]. Furthermore, CD24 is one of the cancer stem cells primarily found in both pancreatic ductal adenocarcinomas and intestinal ductal adenocarcinomas [35, 37, 39]. Furthermore, CD24 expression was strongly associated with poor glandular differentiation, high proliferation and poor prognosis [62]. Analysis of CD24 expression in intestinal and pancreatic neuroendocrine tumors (NETs) revealed that the CD24 has little to no expression in pancreatic and duodenal NETs [71]. A direct link between WNT/β-catenin pathway and CD24 expression was established in the context of PDAC differentiation. The presence of CD24 expression was associated with differentiated tumors [53]. Additionally, studies in human pancreatic cancer xenografts of mice with nonobese diabetic/severe combined immunodeficiency (NOD/SCID) exhibited that the co-expression of CD44+, CD24+, and ESA+ in a subpopulation of malignant cells plays a key role in self-renewal and multidifferentiation potential [46].

9.3.4

CD166

CD166 is described as an activated leukocyte cell adhesion molecule (ALCAM). Elevated expression of CD166 mediates migratory and tumorigenic activity of pa-CSCs and serves as an independent marker for prognosis and tumor relapse [40]. However, expression of ALCAM in the PC lesions is not associated with clinical or pathological data [79]. Subsequent studies delineated that the highly

164

L. Saikrishna et al.

tumorigenic tendency of CD166+ cells was due to the overexpression of TSPAN8 and BST2, while stronger invasive and migratory activities CD166- cells was due to overexpression of BMP7 and Col6A1 [25].

9.3.5

CD326 or EpCAM

Epithelial cell adhesion molecules (EpCAM) are primarily expressed in rapidly proliferating tumors of epithelial origin [83]. EpCAM’s functions include cell signaling, cell-cell adhesion, proliferation, tumorigenesis, differentiation, metastasis and migration [52, 64]. The overexpression of EpCAM was found in many malignant tumors including PC [64, 85]. Compared to the EpCAM- PC cells, EpCAM+ CSCs show an increased tumorigenic potential by a 100-fold [45]. Furthermore, in PC patients, the expression of EpCAM was higher in the cytoplasm but lower in the membranes. However, the EpCAM overexpression suggested a worse survival rate in patients with advanced stage of carcinomas [57]. In contrast to this, Ep-CAM expression was strongly correlated with the suppression of PC, leading to a better prognosis in patients undergoing the resection [2]. In addition, there are no differences in the pathological features in between patients with low and high concentrations/expression of Ep-CAM [2]. Hence, further studies are necessary to validate the clinical significance of EpCAM in PC.

9.3.6

ALDH1

ALDH1 is an intracellular cytosolic enzyme, which plays a primary role in cellular detoxification and differentiation. ALDH1 activity has been associated with both normal stem cells and progenitor cells [59]. The expression of ALDH1 is regulated by various mechanisms including stem cell protein Piwil1 [15], long non-coding RNA HOTTIP [24] and acetylation of lysine 353 [96]. Although increased expression of ALDH-1 in healthy pancreatic tissues disqualifies ALDH1 as a suitable marker for CSCs in humans, (OPTIONAL: investigations show that) ALDH1 is associated with tumorigenic cells in PDAC [17, 42]. Pancreatic CSCs expressing aldehyde dehydrogenase (ADLH) showed an enhanced clonogenic growth and reduced OS of PC patients [68]. Studies in direct xenograft tumors, showed that the cell populations with high ALDH activity alone could enhance tumorigenic potential with efficiency in tumor-initiation [42]. A current study showed that the ALDH1 expression is inversely associated with the expression of 15-Hydroxyprostaglandin dehydrogenase (15-PGDH) and is associated with poor outcomes in PDAC patients [6]. Down-regulation of the expression of ALDH1 significantly reduced the expression of downstream genes and also decreased the growth of PDAC cells [6].

9

Perspectives and Molecular Understanding of Pancreatic Cancer Stem Cells

9.3.7

165

c-Met

c-Met is a receptor of the tyrosine kinase family that serves as a proto-oncogene and is activated by its ligand hepatocyte growth factor (HGF) [58]. As an important marker for pa-CSCs, its signaling induces growth and invasion, and cells that express increased c-Met has tumorigenic potential [47, 78]. c-Met inhibitors given alone or with gemcitabine, reduce tumor growth and prevent the development of metastases [47]. Additionally, inhibiting c-Met via XL184 halts self-renewal capacity in pancreatic CSCs and prevents the development of metastases [31]. Furthermore, targeting the HGF-c-MET pathway inhibits local tumor growth and metastasis [65]. A recent study showed that the HGF-c-MET mediates tumor metabostemness by modulating YAP/HIF-1α [93]. Furthermore, complete suppression of the c-Met signaling pathway with gemcitabine chemotherapy acts as a more effective therapeutic option for PDAC [60]. Some phytochemicals, such as withaferin A and carnosol, are used to suppress c-Met kinase domain to treat in c-Met-dependent cancers [3].

9.3.8

DCLK1

Although the microtubule regulator doublecortin-like kinase 1 (DCLK1) is expressed predominately in invasive and metastatic CSCs, previous studies have noted that it is a normal stem cell marker in the gut [27, 56]. The expression of DCLK1, in recent studies, is described as a biomarker for pa-CSCs [8]. Manipulation DCLK1 and other co-expressed markers (ATAT1, HES1, HEY1, IGF1R, and ABL1) demonstrated a reduction in the clonogenic potential of PDAC cell lines [8]. The expression of DCLK1 was positively correlated with intrahepatic metastasis and poor disease-free survival in patients with hepatocellular carcinoma [20]. The DCLK1 gene was also associated with H3K4me3 and H3K27me3 histone modifications in pancreatic CSCs with invasive and metastatic potential [36]. Overexpression of microRNA-195 inhibited cellular proliferation, cellular migration and invasion of PC cells by targeting DCLK1 [98]. Furthermore, we found that DCLK1 silencing could inactivate/suppress epithelial-mesenchymal transition in cancer cells by downregulating Bmi-1, Snail and Vimentin [50]. Overall these studies revealed that the human pancreatic CSCs showed DCLK1 overexpression and inhibition, suggesting that regulation of DCLK1 expression offers a novel approach in treating pancreatic cancer cells.

9.3.9

CXCR4

Expression C-X-C chemokine receptor type 4 (CXCR4 or CD184) was found on tumor cells of various cancers [9]. Overexpression of CXCR4 has been linked with lung and liver metastases of murine pancreatic cancer [72]. CXCR4 expression has been identified in the stem cell population of various cancers and is linked with

166

L. Saikrishna et al.

aggressiveness and the promotion of metastasis [16, 18, 22, 63]. Furthermore, studies showed that the paCSC with CXCR4+/CD133+ is essential for the metastasis of PC [30]. Additionally, the drug resistance caused by the activation of CXCR4 in pancreatic cancer cells indicates that CXCR4 is a novel therapeutic target for PC therapy [74]. Hypoxia-inducible factor 1 is a strong activator of CXCR4 and CXCL12 expression and the interaction of CXCR4/CXCL12 contributes heavily in cancer cell progression, cellular expansion, invasion, as well as metastasis of PC [73, 89]. Currently CXCR4 antagonists have shown the most striking response in tumor growth delay; thus, a randomized clinical trial is currently under development to study the effectiveness of applying a combination of immune-checkpoint inhibitors with CXCR4 antagonists (NCT03193190).

9.3.10 ABCG2 The ATP-binding cassette subfamily G member 2 (ABCG2) is a cancer stem cell marker in various cancers [76, 88]. The ABCG2 gene is located on chromosome 4q22, which leads to the coding of 655-amino-acid breast cancer resistance protein BCRP [7]. Overexpression of Gastrin and ABCG2 were found in PC cell lines as well as cancer tissues. ABCG2 transporter expression in side-population (SP) cells derived from the human PDAC cells indicates its role in determining SP phenotype [10]. Furthermore, gastrin induced ABCG2 expression by activating NF-κB and thereby alters invasion activity and metastasis in PC [88]. Verapamil, a inhibitor of ABCG2 transporter, increased the responsiveness of PDAC SP cells to the vinca alkaloid vincristine [10]. Furthermore, the drugs that are able to overcome ABCG2 resistance make them an effective therapeutic option against stem-like cancer cell populations in pancreatic tumors [91]. The FL118, an ABCG2 non-substrate anticancer agent, may be able to bypass ABCG2 -mediated drug resistance of human PC [51].

9.4

Conclusions

Pancreatic cancer is a highly aggressive tumor type with dismal prognosis. Current treatments of PC such as surgery, radiotherapy, and chemotherapy are not effective in eliminating the disease. Although some promising drug candidates with antitumor activity against PC have been identified, recent trials of these drugs have shown limited to no benefit. The primary reason for the continual appearance tumors despite radiation or chemotherapy is due to the presence of self-renewing CSCs. Pancreatic cancer CSCs are involved in aggressive, chemo-resistant and metastatic nature of this cancer. This understanding raised the possibility to stratify cancer patients based

9

Perspectives and Molecular Understanding of Pancreatic Cancer Stem Cells

167

on surface markers and the subsequent discovery of novel treatment options for each subtype of PC. The agents targeting the tumor microenvironment and intrinsic signaling pathways targeting CSC include hedgehog, Wnt/ß-catenin, and notch. Conflict of Interest There are no conflicts of interests.

References 1. Adamska A, Domenichini A, Falasca M (2017) Pancreatic ductal adenocarcinoma: current and evolving therapies. Int J Mol Sci 18:1338 2. Akita H, Nagano H, Takeda Y, Eguchi H, Wada H, Kobayashi S, Marubashi S, Tanemura M, Takahashi H, Ohigashi H, Tomita Y, Ishikawa O, Mori M, Doki Y (2011) Ep-CAM is a significant prognostic factor in pancreatic cancer patients by suppressing cell activity. Oncogene 30:3468–3476 3. Aliebrahimi S, Kouhsari SM, Arab SS, Shadboorestan A, Ostad SN (2018) Phytochemicals, withaferin A and carnosol, overcome pancreatic cancer stem cells as c-Met inhibitors. Biomed Pharmacother 106:1527–1536 4. Amundadottir LT (2016) Pancreatic cancer genetics. Int J Biol Sci 12:314–325 5. Ansari D, Tingstedt B, Andersson B, Holmquist F, Sturesson C, Williamsson C, Sasor A, Borg D, Bauden M, Andersson R (2016) Pancreatic cancer: yesterday, today and tomorrow. Future Oncol 12:1929–1946 6. Arima K, Ohmuraya M, Miyake K, Koiwa M, Uchihara T, Izumi D, Gao F, Yonemura A, Bu L, Okabe H, Imai K, Hashimoto D, Baba Y, Chikamoto A, Yamashita YI, Furukawa T, Araki K, Baba H, Ishimoto T (2019) Inhibition of 15-PGDH causes Kras-driven tumor expansion through prostaglandin E2-ALDH1 signaling in the pancreas. Oncogene 38:1211–1224 7. Au A, Aziz Baba A, Goh AS, Wahid Fadilah SA, Teh A, Rosline H, Ankathil R (2014) Association of genotypes and haplotypes of multi-drug transporter genes ABCB1 and ABCG2 with clinical response to imatinib mesylate in chronic myeloid leukemia patients. Biomed Pharmacother 68:343–349 8. Bailey JM, Alsina J, Rasheed ZA, McAllister FM, Fu YY, Plentz R, Zhang H, Pasricha PJ, Bardeesy N, Matsui W, Maitra A, Leach SD (2014) DCLK1 marks a morphologically distinct subpopulation of cells with stem cell properties in preinvasive pancreatic cancer. Gastroenterology 146:245–256 9. Balkwill F (2004) The significance of cancer cell expression of the chemokine receptor CXCR4. Semin Cancer Biol 14:171–179 10. Bhagwandin VJ, Bishop JM, Wright WE, Shay JW (2016) The metastatic potential and chemoresistance of human pancreatic cancer stem cells. PLoS One 11:e0148807 11. Bonnet D, Dick JE (1997) Human acute myeloid leukemia is organized as a hierarchy that originates from a primitive hematopoietic cell. Nat Med 3:730–737 12. Bray F, Ferlay J, Soerjomataram I, Siegel RL, Torre LA, Jemal A (2018) Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA: Cancer J Clin 68:394–424 13. Chatterjee S, Behnam Azad B, Nimmagadda S (2014) The intricate role of CXCR4 in cancer. Adv Cancer Res 124:31–82 14. Chen Z, Che Q, He X, Wang F, Wang H, Zhu M, Sun J, Wan X (2015) Stem cell protein Piwil1 endowed endometrial cancer cells with stem-like properties via inducing epithelialmesenchymal transition. BMC Cancer 15:811 15. Chen F, Roberts NJ, Klein AP (2017) Inherited pancreatic cancer. Chin Clin Oncol 6:58–58

168

L. Saikrishna et al.

16. Costa MJ, Kudaravalli J, Ma JT, Ho WH, Delaria K, Holz C, Stauffer A, Chunyk AG, Zong Q, Blasi E, Buetow B, Tran TT, Lindquist K, Dorywalska M, Rajpal A, Shelton DL, Strop P, Liu SH (2019) Optimal design, anti-tumour efficacy and tolerability of anti-CXCR4 antibody drug conjugates. Sci Rep 9:2443 17. Deng S, Yang X, Lassus H, Liang S, Kaur S, Ye Q, Li C, Wang LP, Roby KF, Orsulic S, Connolly DC, Zhang Y, Montone K, Butzow R, Coukos G, Zhang L (2010) Distinct expression levels and patterns of stem cell marker, aldehyde dehydrogenase isoform 1 (ALDH1), in human epithelial cancers. PLoS One 5:e10277 18. Dimova I, Karthik S, Makanya A, Hlushchuk R, Semela D, Volarevic V, Djonov V (2019) SDF-1/CXCR4 signalling is involved in blood vessel growth and remodelling by intussusception. J Cell Mol Med 23:3916–3926 19. Ding Q, Yoshimitsu M, Kuwahata T, Maeda K, Hayashi T, Obara T, Miyazaki Y, Matsubara S, Natsugoe S, Takao S (2012) Establishment of a highly migratory subclone reveals that CD133 contributes to migration and invasion through epithelial-mesenchymal transition in pancreatic cancer. Hum Cell 25:1–8 20. Fan M, Qian N, Dai G (2017) Expression and prognostic significance of doublecortin-like kinase 1 in patients with hepatocellular carcinoma. Oncol Lett 14:7529–7537 21. Fargeas CA, Florek M, Huttner WB, Corbeil D (2003) Characterization of prominin-2, a new member of the prominin family of pentaspan membrane glycoproteins. J Biol Chem 278:8586–8596 22. Fazi B, Proserpio C, Galardi S, Annesi F, Cola M, Mangiola A, Michienzi A, Ciafre SA (2019) The expression of the chemokine CXCL14 correlates with several aggressive aspects of glioblastoma and promotes key properties of glioblastoma cells. Int J Mol Sci 20. https://doi. org/10.3390/ijms20102496 23. Florek M, Bauer N, Janich P, Wilsch-Braeuninger M, Fargeas CA, Marzesco AM, Ehninger G, Thiele C, Huttner WB, Corbeil D (2007) Prominin-2 is a cholesterol-binding protein associated with apical and basolateral plasmalemmal protrusions in polarized epithelial cells and released into urine. Cell Tissue Res 328:31–47 24. Fu Z, Chen C, Zhou Q, Wang Y, Zhao Y, Zhao X, Li W, Zheng S, Ye H, Wang L, He Z, Lin Q, Li Z, Chen R (2017) LncRNA HOTTIP modulates cancer stem cell properties in human pancreatic cancer by regulating HOXA9. Cancer Lett 410:68–81 25. Fujiwara K, Ohuchida K, Sada M, Horioka K, Ulrich CD 3rd, Shindo K, Ohtsuka T, Takahata S, Mizumoto K, Oda Y, Tanaka M (2014) CD166/ALCAM expression is characteristic of tumorigenicity and invasive and migratory activities of pancreatic cancer cells. PLoS One 9:e107247 26. Genkinger JM, Spiegelman D, Anderson KE, Bernstein L, van den Brandt PA, Calle EE, English DR, Folsom AR, Freudenheim JL, Fuchs CS, Giles GG, Giovannucci E, Horn-Ross PL, Larsson SC, Leitzmann M, Mannisto S, Marshall JR, Miller AB, Patel AV, Rohan TE, Stolzenberg-Solomon RZ, Verhage BA, Virtamo J, Willcox BJ, Wolk A, Ziegler RG, SmithWarner SA (2011) A pooled analysis of 14 cohort studies of anthropometric factors and pancreatic cancer risk. Int J Cancer 129:1708–1717 27. Giannakis M, Stappenbeck TS, Mills JC, Leip DG, Lovett M, Clifton SW, Ippolito JE, Glasscock JI, Arumugam M, Brent MR, Gordon JI (2006) Molecular properties of adult mouse gastric and intestinal epithelial progenitors in their niches. J Biol Chem 281:11292–11300 28. Glumac PM, LeBeau AM (2018) The role of CD133 in cancer: a concise review. Clin Transl Med 7:18–18 29. Hashimoto O, Shimizu K, Semba S, Chiba S, Ku Y, Yokozaki H, Hori Y (2011) Hypoxia induces tumor aggressiveness and the expansion of CD133-positive cells in a hypoxia-inducible factor-1alpha-dependent manner in pancreatic cancer cells. Pathobiol J Immunopathol, Mol Cell Biol 78:181–192

9

Perspectives and Molecular Understanding of Pancreatic Cancer Stem Cells

169

30. Hermann PC, Huber SL, Herrler T, Aicher A, Ellwart JW, Guba M, Bruns CJ, Heeschen C (2007) Distinct populations of cancer stem cells determine tumor growth and metastatic activity in human pancreatic cancer. Cell Stem Cell 1:313–323 31. Herreros-Villanueva M, Zubia-Olascoaga A, Bujanda L (2012) c-Met in pancreatic cancer stem cells: therapeutic implications. World J Gastroenterol 18:5321–5323 32. Hsu CP, Lee LY, Hsu JT, Hsu YP, Wu YT, Wang SY, Yeh CN, Chen TC, Hwang TL (2018) CD44 predicts early recurrence in pancreatic cancer patients undergoing radical surgery. In vivo (Athens, Greece) 32:1533–1540 33. Huynh DL, Koh H, Chandimali N, Zhang JJ, Kim N, Kang TY, Ghosh M, Gera M, Park Y-H, Kwon T, Jeong DK (2019) BRM270 inhibits the proliferation of CD44 positive pancreatic ductal adenocarcinoma cells via downregulation of sonic hedgehog signaling. Evid Based Complement Alternat Med 2019:8 34. Immervoll H, Hoem D, Sakariassen PØ, Steffensen OJ, Molven A (2008) Expression of the “stem cell marker” CD133 in pancreas and pancreatic ductal adenocarcinomas. BMC Cancer 8:48–48 35. Ischenko I, Petrenko O, Hayman MJ (2014) Analysis of the tumor-initiating and metastatic capacity of PDX1-positive cells from the adult pancreas. Proc Natl Acad Sci U S A 111:3466–3471 36. Ito H, Tanaka S, Akiyama Y, Shimada S, Adikrisna R, Matsumura S, Aihara A, Mitsunori Y, Ban D, Ochiai T, Kudo A, Arii S, Yamaoka S, Tanabe M (2016) Dominant expression of DCLK1 in human pancreatic cancer stem cells accelerates tumor invasion and metastasis. PLoS One 11:e0146564–e0146564 37. Jaggupilli A, Elkord E (2012) Significance of CD44 and CD24 as cancer stem cell markers: an enduring ambiguity. Clin Dev Immunol 2012:708036 38. Jarosz M, Sekula W, Rychlik E (2012) Influence of diet and tobacco smoking on pancreatic cancer incidence in Poland in 1960–2008. Gastroenterol Res Pract 2012:682156 39. Jiang W, Sui X, Zhang D, Liu M, Ding M, Shi Y, Deng H (2011) CD24: a novel surface marker for PDX1-positive pancreatic progenitors derived from human embryonic stem cells. Stem Cells (Dayton, Ohio) 29:609–617 40. Kahlert C, Weber H, Mogler C, Bergmann F, Schirmacher P, Kenngott HG, Matterne U, Mollberg N, Rahbari NN, Hinz U, Koch M, Aigner M, Weitz J (2009) Increased expression of ALCAM/CD166 in pancreatic cancer is an independent prognostic marker for poor survival and early tumour relapse. Br J Cancer 101:457–464 41. Kim WT, Ryu CJ (2017) Cancer stem cell surface markers on normal stem cells. BMB Rep 50:285–298 42. Kim MP, Fleming JB, Wang H, Abbruzzese JL, Choi W, Kopetz S, McConkey DJ, Evans DB, Gallick GE (2011) ALDH activity selectively defines an enhanced tumor-initiating cell population relative to CD133 expression in human pancreatic adenocarcinoma. PLoS One 6:e20636 43. Kiuchi S, Ikeshita S, Miyatake Y, Kasahara M (2015) Pancreatic cancer cells express CD44 variant 9 and multidrug resistance protein 1 during mitosis. Exp Mol Pathol 98:41–46 44. Koury J, Zhong L, Hao J (2017) Targeting signaling pathways in Cancer stem cells for Cancer treatment. Stem Cells Int 2017:10 45. Li C, Heidt DG, Dalerba P, Burant CF, Zhang L, Adsay V, Wicha M, Clarke MF, Simeone DM (2007) Identification of pancreatic cancer stem cells. Cancer Res 67:1030–1037 46. Li C, Lee CJ, Simeone DM (2009) Identification of human pancreatic cancer stem cells. Methods Mol Biol 568:161–173 47. Li C, Wu JJ, Hynes M, Dosch J, Sarkar B, Welling TH, Pasca di Magliano M, Simeone DM (2011) c-Met is a marker of pancreatic cancer stem cells and therapeutic target. Gastroenterology 141:2218–2227.e5 48. Li Z, Chen K, Jiang P, Zhang X, Li X, Li Z (2014) CD44v/CD44s expression patterns are associated with the survival of pancreatic carcinoma patients. Diagn Pathol 9:79 49. Li X-P, Zhang X-W, Zheng L-Z, Guo W-J (2015) Expression of CD44 in pancreatic cancer and its significance. Int J Clin Exp Pathol 8:6724–6731

170

L. Saikrishna et al.

50. Li J, Wang Y, Ge J, Li W, Yin L, Zhao Z, Liu S, Qin H, Yang J, Wang L, Ni B, Liu Y, Wang H (2018) Doublecortin-like kinase 1 (DCLK1) regulates B cell-specific Moloney murine leukemia virus insertion site 1 (Bmi-1) and is associated with metastasis and prognosis in pancreatic Cancer. Cell Physiol Biochem 51:262–277 51. Ling X, Wu W, Fan C, Xu C, Liao J, Rich LJ, Huang R-Y, Repasky EA, Wang X, Li F (2018) An ABCG2 non-substrate anticancer agent FL118 targets drug-resistant cancer stem-like cells and overcomes treatment resistance of human pancreatic cancer. J Exp Clin Cancer Res: CR 37:240–240 52. Litvinov SV, Velders MP, Bakker HA, Fleuren GJ, Warnaar SO (1994) Ep-CAM: a human epithelial antigen is a homophilic cell-cell adhesion molecule. J Cell Biol 125:437–446 53. Lubeseder-Martellato C, Hidalgo-Sastre A, Hartmann C, Alexandrow K, KamyabiMoghaddam Z, Sipos B, Wirth M, Neff F, Reichert M, Heid I, Schneider G, Braren R, Schmid RM, Siveke JT (2016) Membranous CD24 drives the epithelial phenotype of pancreatic cancer. Oncotarget 7:49156–49168 54. Maccalli C, Volonte A, Cimminiello C, Parmiani G (2014) Immunology of cancer stem cells in solid tumours. A review. Eur J Cancer 50:649–655 55. Maisonneuve P, Lowenfels AB (2014) Risk factors for pancreatic cancer: a summary review of meta-analytical studies. Int J Epidemiol 44:186–198 56. May R, Riehl TE, Hunt C, Sureban SM, Anant S, Houchen CW (2008) Identification of a novel putative gastrointestinal stem cell and adenoma stem cell marker, doublecortin and CaM kinaselike-1, following radiation injury and in adenomatous polyposis coli/multiple intestinal neoplasia mice. Stem Cells (Dayton, Ohio) 26:630–637 57. Meng Y, Xu B-Q, Fu Z-G, Wu B, Xu B, Chen Z-N, Li L (2015) Cytoplasmic EpCAM overexpression is associated with favorable clinical outcomes in pancreatic cancer patients with Hepatitis B virus negative infection. Int J Clin Exp Med 8:22204–22216 58. Michieli P, Mazzone M, Basilico C, Cavassa S, Sottile A, Naldini L, Comoglio PM (2004) Targeting the tumor and its microenvironment by a dual-function decoy Met receptor. Cancer Cell 6:61–73 59. Moreb J, Schweder M, Suresh A, Zucali JR (1996) Overexpression of the human aldehyde dehydrogenase class I results in increased resistance to 4-hydroperoxycyclophosphamide. Cancer Gene Ther 3:24–30 60. Noguchi K, Konno M, Eguchi H, Kawamoto K, Mukai R, Nishida N, Koseki J, Wada H, Akita H, Satoh T, Marubashi S, Nagano H, Doki Y, Mori M, Ishii H (2018) c-Met affects gemcitabine resistance during carcinogenesis in a mouse model of pancreatic cancer. Oncol Lett 16:1892–1898 61. Nomura A, Banerjee S, Chugh R, Dudeja V, Yamamoto M, Vickers SM, Saluja AK (2015) CD133 initiates tumors, induces epithelial-mesenchymal transition and increases metastasis in pancreatic cancer. Oncotarget 6:8313–8322 62. Ohara Y, Oda T, Sugano M, Hashimoto S, Enomoto T, Yamada K, Akashi Y, Miyamoto R, Kobayashi A, Fukunaga K, Morishita Y, Ohkohchi N (2013) Histological and prognostic importance of CD44(+) /CD24(+) /EpCAM(+) expression in clinical pancreatic cancer. Cancer Sci 104:1127–1134 63. Pan WL, Wang Y, Hao Y, Wong JH, Chan WC, Wan DC, Ng TB (2018) Overexpression of CXCR4 synergizes with LL-37 in the metastasis of breast cancer cells. Biochim Biophys Acta Mol basis Dis 1864:3837–3846 64. Patriarca C, Macchi RM, Marschner AK, Mellstedt H (2012) Epithelial cell adhesion molecule expression (CD326) in cancer: a short review. Cancer Treat Rev 38:68–75 65. Pothula SP, Xu Z, Goldstein D, Biankin AV, Pirola RC, Wilson JS, Apte MV (2016) Hepatocyte growth factor inhibition: a novel therapeutic approach in pancreatic cancer. Br J Cancer 114:269 66. Prasetyanti PR, Medema JP (2017) Intra-tumor heterogeneity from a cancer stem cell perspective. Mol Cancer 16:41–41

9

Perspectives and Molecular Understanding of Pancreatic Cancer Stem Cells

171

67. Rall CJ, Rustgi AK (1995) CD44 isoform expression in primary and metastatic pancreatic adenocarcinoma. Cancer Res 55:1831–1835 68. Rasheed ZA, Yang J, Wang Q, Kowalski J, Freed I, Murter C, Hong SM, Koorstra JB, Rajeshkumar NV, He X, Goggins M, Iacobuzio-Donahue C, Berman DM, Laheru D, Jimeno A, Hidalgo M, Maitra A, Matsui W (2010) Prognostic significance of tumorigenic cells with mesenchymal features in pancreatic adenocarcinoma. J Natl Cancer Inst 102:340–351 69. Rawla P, Sunkara T, Gaduputi V (2019) Epidemiology of pancreatic cancer: global trends, etiology and risk factors. World J Oncol 10:10–27 70. Sakaue T, Koga H, Iwamoto H, Nakamura T, Ikezono Y, Abe M, Wada F, Masuda A, Tanaka T, Fukahori M, Ushijima T, Mihara Y, Naitou Y, Okabe Y, Kakuma T, Ohta K, Nakamura KI, Torimura T (2019) Glycosylation of ascites-derived exosomal CD133: a potential prognostic biomarker in patients with advanced pancreatic cancer. Med Mol Morphol. https://doi.org/10.1007/s00795-019-00218-5 71. Salaria S, Means A, Revetta F, Idrees K, Liu E, Shi C (2015) Expression of CD24, a stem cell marker, in pancreatic and small intestinal neuroendocrine tumors. Am J Clin Pathol 144:642–648 72. Saur D, Seidler B, Schneider G, Algul H, Beck R, Senekowitsch-Schmidtke R, Schwaiger M, Schmid RM (2005) CXCR4 expression increases liver and lung metastasis in a mouse model of pancreatic cancer. Gastroenterology 129:1237–1250 73. Shakir M, Tang D, Zeh HJ, Tang SW, Anderson CJ, Bahary N, Lotze MT (2015) The chemokine receptors CXCR4/CXCR7 and their primary heterodimeric ligands CXCL12 and CXCL12/high mobility group box 1 in pancreatic cancer growth and development: finding flow. Pancreas 44:528–534 74. Singh S, Srivastava SK, Bhardwaj A, Owen LB, Singh AP (2010) CXCL12-CXCR4 signalling axis confers gemcitabine resistance to pancreatic cancer cells: a novel target for therapy. Br J Cancer 103:1671–1679 75. Singh S, Arcaroli JJ, Orlicky DJ, Chen Y, Messersmith WA, Bagby S, Purkey A, Quackenbush KS, Thompson DC, Vasiliou V (2016) Aldehyde dehydrogenase 1B1 as a modulator of pancreatic adenocarcinoma. Pancreas 45:117–122 76. Sobek KM, Cummings JL, Bacich DJ, O’Keefe DS (2017) Contrasting roles of the ABCG2 Q141K variant in prostate cancer. Exp Cell Res 354:40–47 77. Sugahara KN, Hirata T, Hayasaka H, Stern R, Murai T, Miyasaka M (2006) Tumor cells enhance their own CD44 cleavage and motility by generating hyaluronan fragments. J Biol Chem 281:5861–5868 78. Suzuki A, Nakauchi H, Taniguchi H (2004) Prospective isolation of multipotent pancreatic progenitors using flow-cytometric cell sorting. Diabetes 53:2143–2152 79. Tachezy M, Zander H, Marx AH, Stahl PR, Gebauer F, Izbicki JR, Bockhorn M (2012) ALCAM (CD166) expression and serum levels in pancreatic cancer. PLoS One 7:e39018– e39018 80. Tirino V, Desiderio V, Paino F, De Rosa A, Papaccio F, La Noce M, Laino L, De Francesco F, Papaccio G (2013) Cancer stem cells in solid tumors: an overview and new approaches for their isolation and characterization. FASEB J 27:13–24 81. Todaro M, Francipane MG, Medema JP, Stassi G (2010) Colon cancer stem cells: promise of targeted therapy. Gastroenterology 138:2151–2162 82. Tomita H, Tanaka K, Tanaka T, Hara A (2016) Aldehyde dehydrogenase 1A1 in stem cells and cancer. Oncotarget 7:11018–11032 83. Trzpis M, McLaughlin PMJ, de Leij LMFH, Harmsen MC (2007) Epithelial cell adhesion molecule: more than a carcinoma marker and adhesion molecule. Am J Pathol 171:386–395 84. Ulasov IV, Nandi S, Dey M, Sonabend AM, Lesniak MS (2011) Inhibition of sonic hedgehog and notch pathways enhances sensitivity of CD133(+) glioma stem cells to temozolomide therapy. Mol Med (Cambridge, Mass.) 17:103–112 85. van der Gun BT, Melchers LJ, Ruiters MH, de Leij LF, McLaughlin PM, Rots MG (2010) EpCAM in carcinogenesis: the good, the bad or the ugly. Carcinogenesis 31:1913–1921 86. Vander Linden C, Corbet C (2019) Therapeutic targeting of cancer stem cells: integrating and exploiting the acidic niche. Front Oncol 9:159

172

L. Saikrishna et al.

87. Vassalli G (2019) Aldehyde dehydrogenases: not just markers, but functional regulators of stem cells. Stem Cells Int 2019:15 88. Wang J, Xin B, Wang H, He X, Wei W, Zhang T, Shen X (2016) Gastrin regulates ABCG2 to promote the migration, invasion and side populations in pancreatic cancer cells via activation of NF-kappaB signaling. Exp Cell Res 346:74–84 89. Wang X, Cao Y, Zhang S, Chen Z, Fan L, Shen X, Zhou S, Chen D (2017) Stem cell autocrine CXCL12/CXCR4 stimulates invasion and metastasis of esophageal cancer. Oncotarget 8:36149–36160 90. Wang RQ, Geng J, Sheng WJ, Liu XJ, Jiang M, Zhen YS (2019) The ionophore antibiotic gramicidin A inhibits pancreatic cancer stem cells associated with CD47 down-regulation. Cancer Cell Int 19:145 91. Westover D, Li F (2015) New trends for overcoming ABCG2/BCRP-mediated resistance to cancer therapies. J Exp Clin Cancer Res: CR 34:159 92. Yamamoto T, Yagi S, Kinoshita H, Sakamoto Y, Okada K, Uryuhara K, Morimoto T, Kaihara S, Hosotani R (2015) Long-term survival after resection of pancreatic cancer: a single-center retrospective analysis. World J Gastroenterol 21:262–268 93. Yan B, Jiang Z, Cheng L, Chen K, Zhou C, Sun L, Qian W, Li J, Cao J, Xu Q, Ma Q, Lei J (2018) Paracrine HGF/c-MET enhances the stem cell-like potential and glycolysis of pancreatic cancer cells via activation of YAP/HIF-1α. Exp Cell Res 371:63–71 94. Yuan Y, Yang Z, Miao X, Li D, Liu Z, Zou Q (2015) The clinical significance of FRAT1 and ABCG2 expression in pancreatic ductal adenocarcinoma. Tumour Biol 36:9961–9968 95. Zhan T, Rindtorff N, Boutros M (2017) Wnt signaling in cancer. Oncogene 36:1461–1473 96. Zhao D, Mo Y, Li MT, Zou SW, Cheng ZL, Sun YP, Xiong Y, Guan KL, Lei QY (2014) NOTCH-induced aldehyde dehydrogenase 1A1 deacetylation promotes breast cancer stem cells. J Clin Invest 124:5453–5465 97. Zhao S, Chen C, Chang K, Karnad A, Jagirdar J, Kumar AP, Freeman JW (2016) CD44 expression level and isoform contributes to pancreatic Cancer cell plasticity, invasiveness, and response to therapy. Clin Cancer Res 22:5592–5604 98. Zhou B, Sun C, Hu X, Zhan H, Zou H, Feng Y, Qiu F, Zhang S, Wu L, Zhang B (2017) MicroRNA-195 suppresses the progression of pancreatic cancer by targeting DCLK1. Cell Physiol Biochem 44:1867–1881 99. Zhu L, Finkelstein D, Gao C, Shi L, Wang Y, Lopez-Terrada D, Wang K, Utley S, Pounds S, Neale G, Ellison D, Onar-Thomas A, Gilbertson RJ (2016) Multi-organ mapping of cancer risk. Cell 166:1132–1146.e7

The Role of Hypoxia Inducible Factor-1a in Pancreatic Cancer and Diabetes Mellitus

10

Saimila Momin and Ganji Purnachandra Nagaraju

Abstract

Hypoxia inducible factor-1α, abbreviated as HIF-1α, is a protein which is encoded by the HIF1A gene and is primarily responsible for the modulation of cellular feedback regarding hypoxia, which results from the lack of proper delivery of oxygen to cells. When HIF-1α is unregulated or over-expressed due to hypoxia, it can lead to irregular cellular metabolism and pancreatic cancer (PC) progression. There seems to be a strong, positive correlation between up regulation of HIF-1α and cancers, specifically PC. Understanding the regulation, expression, and interaction of HIF-1α will allows us to further understand the role this pathway can play in deadly diseases such as PC. In this chapter we will explore the role of the HIF-1α pathway in PC, and in metabolic disorders, such as diabetes mellitus, and studies that have been conducted in order to further holistically understand the role of HIF-1α. Keywords

Hypoxia inducible factor 1-α · Pancreatic cancer · Diabetes mellitus · Diseases · Signaling pathway

Abbreviations ARNT bHLH-PAS HIF-1α HSP90

aryl hydrocarbon nuclear translator Basic helix-loop-helix – per-arnt-sim domain Hypoxia inducible factor-1α heat shock protein 90

S. Momin · G. P. Nagaraju (*) Department of Hematology and Medical Oncology, Winship Cancer Institute, School of Medicine, Emory University, Atlanta, GA, USA e-mail: [email protected] # Springer Nature Singapore Pte Ltd. 2019 G. P. Nagaraju, A. BM Reddy (eds.), Exploring Pancreatic Metabolism and Malignancy, https://doi.org/10.1007/978-981-32-9393-9_10

173

174

S. Momin and G. P. Nagaraju

KRAS NAD+/NADH PC pVHL RACK1 SWI-SNF TAD TGF-β

10.1

Kirsten rat sarcoma viral oncogene homolog Nicotinamide adenine dinuclotide Pancreatic cancer von Hippel-Lindau protein complex Receptor for activated C kinase 1 Switch/sucrose non-fermentable topologically associated domain Transforming growth factor beta

Introduction

10.1.1 Pancreatic Cancer Pancreas is a gastrointestinal organ that is primarily responsible for releasing hormones and enzymes, also known as digestive juices, in order to aid in the digestion of foods [1]. The pancreas has both an endocrine and exocrine function because it secretes and releases digestive enzymes into both the bloodstream and into ducts, using its exocrine glands [2]. The pancreas primarily releases the following enzymes: lipase to digest fats, trypsin and chymotrypsin to digest proteins, and amylase to digest carbohydrates [2]. Overall, in the digestive system, digestive juices are released initially into the pancreatic ducts and then the duodenum of the small intestine in order to digest fats, carbohydrates, and proteins, with the additional help of bile, a digestive juice produced in the liver [2]. As mentioned above, along with the exocrine function, the pancreas also has a endocrine function which plays a key role in blood glucose regulation [2]. Islet cells of the pancreas generate and release insulin and glucagon in order to regulate blood glucose levels [3]. Insulin is released to lower blood glucose levels and glucagon is released to increase blood glucose levels. Maintaining blood glucose levels is critical as organ function and overall metabolic function heavily depends on it [3]. Abnormal function of the pancreas can lead to deadly diseases such as pancreatic cancer and diabetes mellitus [4]. Pancreatic cancer (PC) results when cells within this gland organ begin to abnormally divide and grow due to mutations [5]. These mutations arise due to either or both genetic and environmental factors, which include diets, obesity, and substance use [5]. Additionally, even diabetes can increase the likelihood of developing PC [5]. Even though PC is considered to be a deadly and unresolved medical disease, there are current investigations being conducted in order to understand in the underlying pathophysiology and potential treatment plans to tackle the disease. The current treatment plan for PC includes a combination of chemotherapy, interventional radiology, and radiation therapy, which is to attack tumors and decrease their presence [6]. However, there are ongoing investigations studying the molecular mechanisms in order to attack the tumor and overall progression of cancer at its core. These investigations include but are not limited to: AKT inhibitors [7],

10

The Role of Hypoxia Inducible Factor-1a in Pancreatic Cancer and. . .

175

Hypoxic environment

Growth factors

HIF-1β

PI3K HIF-1α

P300

HIF-1α

HIF-1β

HRE

Cell metabolism Cell division Cell growth Cell survival

Nucleus Fig. 10.1 The pathway begins with growth factors that ultimately leads to HIF-1α and HIF-1β subunits with p300 binding to HRE to lead to metabolism, cell survival and cellular proliferation in a hypoxic environment PIK3: phosphoinositide 3-kinase (enzymes); HIF-1α/β: hypoxia inducible factor-1 alpha/beta; HRE: hypoxia response element Masoud GN, Li W (2015) HIF-1α pathway: role, regulation and intervention for cancer therapy. Acta Pharm Sin B 5(5): 378–389

integration of albumin-bound paclitaxel (nab-paclitaxel) plus gemcitabine [8], and other studies in chemotherapy and radiation therapy. Additionally, even though the pathophysiology of PC is still not fully understood, ongoing integrated genomic analyses have revealed 32 mutated genes and 10 pathways that are strongly associated with the increased likelihood of PC [9]. These include KRAS, TGF-β, WNT, NOTCH, ROBO/SLIT signaling, G1/Stransition, SWI-SNF, chromatin alteration, DNA repair function and RNA generation. Along with these identified mutated genes and molecular pathways, hypoxia-inducible factor-1α (HIF-1α) has also been considered to be strong associated with PC along with other cancers [10]. Additionally, by understanding HIF-1α, specifically its role and function, interaction in the body, and how it can be controlled, we can get one step closer to potentially learning how to treat PC and even other cancers that are affected and/or associated with HIF-1α (Fig. 10.1).

10.1.2 Diabetes Mellitus Diabetes is a metabolic disorder that negatively affects how the body digests food and obtains energy for healthy organ and cellular function [11]. Glucose is the primary source of energy in the body and is obtained by foods digested in the

176

S. Momin and G. P. Nagaraju

body and used by all cells in order to properly function. Diabetes generally develops due to the abnormal production of insulin in the body by the pancreas [11]. Insulin is required in order to transport glucose into the cells, which is then converted into energy [12]. Individuals diagnosed with diabetes generate little to no insulin and therefore cellular function decreases and the level of blood glucose increases [13]. In order to compensate for the increased blood glucose levels, the body filters the blood and eliminates the excess glucose. Diabetes can be broken down into two types, type 1 diabetes and type 2 diabetes. Type 1 diabetes is described as the following [14]: a chronic condition in which the islet cells of the pancreas are unable to generate insulin because the immune system is attacking the islet cells of the pancreas. On the other hand, type 2 diabetes is also a chronic condition but it is not an auto-immune disease and develops gradually in which the hyperglycemia leads to insulin resistance. Contributing factors include genetics and environmental factors, such as obesity, unhealthy and high glucose and/or carbohydrate diets, and inactive lifestyle.

10.1.3 Hypoxia Inducible Factor 1 (HIF-1) Hypoxia inducible factor 1 is a heterodimeric transcription factor, which is constructed by two key subunits, an α and a β subunit, which belong to the bHLH-PAS protein family [15]. HIF-1α is an oxygen dependent or sensitive unit which is primarily associated with hypoxia; in other words, when the body undergoes a hypoxia, the α subunit is activated [16]. On the other hand, HIF-1β, which is referred to the aryl hydrocarbon nuclear translator or ARNT acts as a ligand which binds to the aryl hydrocarbon receptor or AhR [15]; upon binding of these two entities, it leads to the translocation to the nucleus. The subunits have domains such as, b, HLH, PAS and TAD, are able to self-regulate their expression [15]. As mentioned before, the HIF-1α plays a key role in aerobic metabolism in order to modulate hemostasis, primarily in cellular oxygen levels. In mammalian bodies, under normal conditions, HIF-1α is transcribed and expressed continuously but eliminated via degradation with a half-life of approximately 5 min [17]. However, when the body undergoes a state of hypoxia, multiple pathways are activated in order to maintain the stability of the α subunit; functions that become activated include: hydroxylation, acetylation, phosphorylation reactions, and ubiquitination, which is primarily mediated by the von Hippel-Lindau protein [15, 18]. Hypoxia can cause from a catastrophe at any step in the transport of oxygen to cells. That leads to decreased limited pressures of oxygen, difficulties with distribution of oxygen in the lungs, inadequate oxygen in hemoglobin [19]. In this case hypoxia, more specifically reaction oxygen species, including nitric oxide, induces the α subunit of the HIF-1. Overall, HIF-1 is responsible for oxygen delivery to cells through the process of angiogenesis [20], which is the process of generating new blood vessels from pre-existing blood vessels and adapting to a hypoxic environment through the process of glycolysis, the first metabolic step to ultimately convert glucose into energy. PC growth and metabolic disorders are strongly correlated to the HIF-1

10

The Role of Hypoxia Inducible Factor-1a in Pancreatic Cancer and. . .

177

expression, specifically over-expression of HIF-1α [21, 22]. Over-expression of this transcription factor leads to PC progression, as tumor cells are able to activate survival pathway to adjust and adapt to the hypoxic environment [18, 23]. Understanding the molecular mechanism and the connection to the pathophysiology of PC and diabetes mellitus can potentially lead to advance treatments for such diseases and others.

10.2

Hypoxia Inducible Factor 1 (HIF-1): Molecular Pathway, Interactions and Regulation

10.2.1 HIF-1 and Pancreatic Cancer As discussed, HIF-1 is a transcription factor that, under hypoxic conditions, primarily modulates the transcription of other genes in order to regulate oxygen levels within the body [24]. The primary subunit which plays an essential role in the control of oxygen levels is the α subunit. The α subunit is modulated by the O2-dependent hydroxylation of proline residue 402, 564, or both, by prolyl hydroxylase domain protein 2 [25]. This leads to the binding and activation of the von Hippel-Lindau protein [25]. Upon binding other processes such as proteasomal degradation and ubiquitination begin to follow, along with oxygen-dependent hydroxylation of asparagine residue 803 [18]. These series of reactions lead to the use of oxygen and α-ketoglutarate to produce carbon dioxide and succinate. As a result of the low levels of oxygen and high levels of products of the tricarboxylic acid cycle, which include isocitrate, succinate, and fumarate, hydroxylase activity is halted [25]. This leads to competitive binding of HIF-1α between the receptor for C kinase 1 or RACK1 and heat shock protein 90 (HSP90). The receptor for C kinase 1 modulates proteasomal degradation and ubiquitination [25]. Overall, under hypoxic conditions, HIF-1α transactivation domains interact with co-activators and leads to the production of reactive oxygen species [24, 25]. Under normal conditions, HIF-1α is hydroxylated by HIF-1α-specific prolyl hydroxylases, which are sensitive to oxygen levels [24]. The process of hydroxylation leads to ubiquitination and proteasomal degradation via the von Hippel-Lindau protein (pVHL) complex. Additionally, HIF-1α is also a substrate a hydroxylase factor that inhibits HIF-1α known as asparaginyl hydroxylase [25]. Since this hydroxylase factor is also sensitive to oxygen levels, it has the ability to interrupt the interaction between HIF-1α transactivation domains and co-activators, which ultimately results in the inhibition of HIF transcriptional activation [25]. There is a strong correlation between the increased concentration of HIF-1α and generation of tumor and PC progression. It is highly suggested that tumor cells are able to adapt to hypoxic conditions and resist apoptosis due to this over-expression of HIF-1α subunit. An experimental study revealed that 13 of 19 diverse tumor types, HIF-1α was severely over-expressed and was also strongly correlated to cellular proliferation [26]. When the body is under hypoxic conditions, processes such as glycolysis and neovascularization are dramatically increased, leading to the

178

S. Momin and G. P. Nagaraju

production of tumors and metastasis [26]. The HIF-1 pathway has the ability to transcribe genes that are responsible for transportation of glucose and such activity is dominated by the HIF-1α unit [26]. Overexpression of HIF-1α leads to impairment in the overall HIF-1 pathway [27], which ultimately leads to tumors cells being able to adapt to such hypoxic conditions and allowing them to thrive in such conditions; whereas, normal cells are unable to properly function in hypoxic conditions. The strong connection between the HIF-1 pathway and HIF-1α subunit to PC progression serves as an attractive topic to explore for therapeutic options [28]. Theoretically, decreasing or down-regulating the HIF-1 pathway can stop tumor progression. Investigations are currently exploring how activating hydroxylase can decrease the activity of the HIF-1 pathway [23]. Other studies are exploring how proteins such as co-activators of the HIF-1α can be altered or deleted in order to prevent transcription [23]. Further exploration and investigation into the manipulation and alteration of such small molecules will lead to decrease of HIF-1 activity and may suppress tumor progression in individuals diagnosed with PC.

10.2.2 HIF-1 and Diabetes Individuals diagnosed with diabetes are often associated with another condition known as pseudohypoxia. Pseudohypoxia is described as having an increased concentration of NADH compared to NAD+ which is primarily because of the increased blood glucose levels (through the polyol pathway [29, 30]. Specifically, studies have shown that diabetic tissues show increased levels of pimonidazole, which is a strong marker for hypoxia, suggesting that diabetic individuals have high levels of hypoxia [30–32]. Additionally, there was also increased expression of HIF-1α [30]. Recall, that the structure of HIF-1α can be easily stabilized by hypoxia and the process of glycolysis. However, individuals with diabetes have decreased transcriptional activity of HIF-1 due to an impairment of transactivation of HIF-1 expression, primarily because of the p300 co-activator [30]. Please note that this impairment does not negatively affect the HIF-1α structural stability. Furthermore, increased glucose levels have the ability to lead to signal transduction (HIF-1 mediated) by ChREBP, a binding protein [30]. Ultimately, experiments have shown that high levels of glucose lead to inactivation and destruction of the HIF-1α subunit and overall impairment of the HIF-1 pathway. Ultimately, cells in the body are unable to properly adapt and respond to the hypoxic conditions, which can lead to immediate cellular dysfunction and other long-term, negative consequences.

10.3

Current Experiments and Investigations

Recently, researchers used triptolide, which is a diterpenoid epoxide extracted from Tripterygium wilfordii and is well-known for its anti-tumor properties, and treated subjects with it to observe the effects it may have on the HIF-1α protein [33]. Upon

10

The Role of Hypoxia Inducible Factor-1a in Pancreatic Cancer and. . .

179

exposure to triptolide, the concentration of HIF-1α increased and the hypoxic response decreased [33]. Additionally, HIF-1 activity was also decreased due to the decrease of the co-activator p300 caused by triptolide [33, 34]. And additionally, triptolide also decreased oncogenic signaling pathways [33]. This substance definitely stands as a strong treatment option because it has the ability to manipulate the HIF-1α at the molecular level and decrease the hypoxic environment. Another study showed that the deletion of pancreas-specific HIF-1α leads to PC and increased concentration of B lymphocytes [35]. However, deletion of the B lymphocytes led to the decrease of PC progression [35]. This suggests that HIF-1α may have a strong and necessary role in preventing tumor progression. Specifically, in this this study, researchers used a KrasG12D-driven murine model and tumors from humans to show that hypoxia and the structural stabilization of HIF-1α are present during the preliminary stages of PC [35]. When the HIF-1α is deleted in pancreas-specific cells, neoplasia is significantly increased, suggesting that HIF-1α plays a key role in regulating and possibly suppressing PC progression.

10.4

Conclusion

Cancer is one of the most intricate mysteries known to the human body, there are many molecular mechanisms and cellular processes that govern and modulate optimal function and the HIF-1α pathway is a particularly relevant mechanism that plays a major, yet complex role in PC and diabetes. Even though the subunit, HIF-1α’s role is understood, it is critical to understand how this subunit interacts with other molecules and if it plays additional roles in other body mechanisms. Possible experimental studies include: exploring interactions of HIF-1α with other organs and other molecular and cellular functions and pathways, identify contributing factors in the pathway that can inhibit or activate HIF-1α pathway, manipulation of the pathway and integrate such changes in live subjects, exploring activation of hydroxylases, which are able to decrease overall HIF-1 activity, targeting the HIF-1 complex for eventual degradation, studying the interaction of the p300, the co-activator of HIF-1α, and observing specific molecules that can serve as inhibitors. Once the implications and side effects of these relatively novice treatment plans and preliminary research studies are properly understood, then these treatment options can undergo clinical trials and studies with human patients, which can ultimately lead to permanent solutions to PC. By further examining this mechanism and studying specifically how the pathway can be manipulated may lead to new therapeutic treatments benefitting those diagnosed with cancers and metabolic disorders. Even though, thus far, as a scientific community, we have not been able to successfully manipulate and alter a transcription factor in a live subject, this method still remains to be a strong therapeutic option; if achieved and applied properly, it could potentially treat PC and other cancers.

180

S. Momin and G. P. Nagaraju

References 1. Grossman M (1950) Gastrointestinal hormones. Physiol Rev 30(1):33–90 2. Henderson J, Daniel P, Fraser P (1981) The pancreas as a single organ: the influence of the endocrine upon the exocrine part of the gland. Gut 22(2):158 3. Ferber S, Halkin A, Cohen H, Ber I, Einav Y, Goldberg I, Barshack I, Seijffers R, Kopolovic J, Kaiser N (2000) Pancreatic and duodenal homeobox gene 1 induces expression of insulin genes in liver and ameliorates streptozotocin-induced hyperglycemia. Nat Med 6(5):568 4. Chari ST, Leibson CL, Rabe KG, Timmons LJ, Ransom J, De Andrade M, Petersen GM (2008) Pancreatic cancer–associated diabetes mellitus: prevalence and temporal association with diagnosis of cancer. Gastroenterology 134(1):95–101 5. Kamisawa T, Wood LD, Itoi T, Takaori K (2016) Pancreatic cancer. Lancet 388(10039):73–85 6. Kleeff J, Korc M, Apte M, La Vecchia C, Johnson CD, Biankin AV, Neale RE, Tempero M, Tuveson DA, Hruban RH (2016) Pancreatic cancer. Nat Rev Dis Primers 2:16022 7. Nitulescu GM, Margina D, Juzenas P, Peng Q, Olaru OT, Saloustros E, Fenga C, Spandidos DΑ, Libra M, Tsatsakis AM (2016) Akt inhibitors in cancer treatment: the long journey from drug discovery to clinical use. Int J Oncol 48(3):869–885 8. Von Hoff DD, Ervin T, Arena FP, Chiorean EG, Infante J, Moore M, Seay T, Tjulandin SA, Ma WW, Saleh MN (2013) Increased survival in pancreatic cancer with nab-paclitaxel plus gemcitabine. N Engl J Med 369(18):1691–1703 9. Bailey P, Chang DK, Nones K, Johns AL, Patch A-M, Gingras M-C, Miller DK, Christ AN, Bruxner TJ, Quinn MC (2016) Genomic analyses identify molecular subtypes of pancreatic cancer. Nature 531(7592):47 10. Wang Z, Dabrosin C, Yin X, Fuster MM, Arreola A, Rathmell WK, Generali D, Nagaraju GP, El-Rayes B, Ribatti D (2015) Broad targeting of angiogenesis for cancer prevention and therapy. Semin Cancer Biol, Elsevier 35:S224–S243 11. Nadal A (2019) Actions of endocrine disrupting chemicals on pancreatic beta cells and risk of diabetes mellitus, 21st European Congress of endocrinology. BioScientifica 63:S7.2 12. Sonksen P, Sonksen J (2000) Insulin: understanding its action in health and disease. Br J Anaesth 85(1):69–79 13. Rains JL, Jain SK (2011) Oxidative stress, insulin signaling, and diabetes. Free Radic Biol Med 50(5):567–575 14. American Diabetes Association (2013) Diagnosis and classification of diabetes mellitus. Diabetes Care 36(Supplement 1):S67–S74 15. Masoud GN, Li W (2015) HIF-1α pathway: role, regulation and intervention for cancer therapy. Acta Pharm Sin B 5(5):378–389 16. Harris AL (2002) Hypoxia—a key regulatory factor in tumour growth. Nat Rev Cancer 2(1):38 17. Anastasiadis AG, Ghafar MA, Salomon L, Vacherot F, Benedit P, Chen M-W, Shabsigh A, Burchardt M, Chopin DK, Shabsigh RJ, oncology c (2002) Human hormone-refractory prostate cancers can harbor mutations in the O 2-dependent degradation domain of hypoxia inducible factor-1α (HIF-1α). J Cancer Res Clin Oncol 128(7):358–362 18. Semenza GL (2001) HIF-1 and mechanisms of hypoxia sensing. Curr Opin Cell Biol 13 (2):167–171 19. Overgaard J, Horsman MR (1996) Modification of hypoxia-induced radioresistance in tumors by the use of oxygen and sensitizers. Semin Radiat Oncol, Elsevier 6:10–21 20. Maxwell PH, Ratcliffe PJ (2002) Oxygen sensors and angiogenesis. Semin Cell Dev Biol, Elsevier 13:29–37 21. Semenza GL (2002) HIF-1 and tumor progression: pathophysiology and therapeutics. Trends Mol Med 8(4):S62–S67 22. Shibaji T, Nagao M, Ikeda N, Kanehiro H, Hisanaga M, Ko S, Fukumoto A, Nakajima Y (2003) Prognostic significance of HIF-1 alpha overexpression in human pancreatic cancer. Anticancer Res 23(6C):4721–4727

10

The Role of Hypoxia Inducible Factor-1a in Pancreatic Cancer and. . .

181

23. Ziello JE, Jovin IS, Huang Y (2007) Hypoxia-Inducible Factor (HIF)-1 regulatory pathway and its potential for therapeutic intervention in malignancy and ischemia. Yale J Biol Med 80(2):51 24. Semenza GL, Shimoda LA, Prabhakar NR (2006) Regulation of gene expression by HIF-1, Novartis Foundation Symposium, Wiley Online Library. John Wiley and Sons, New York, p 2 25. Semenza GL (2007) Hypoxia-inducible factor 1 (HIF-1) pathway. Sci STKE 2007(407):cm8– cm8 26. Zhong H, De Marzo AM, Laughner E, Lim M, Hilton DA, Zagzag D, Buechler P, Isaacs WB, Semenza GL, Simons JW (1999) Overexpression of hypoxia-inducible factor 1α in common human cancers and their metastases. Cancer Res 59(22):5830–5835 27. Ke Q, Costa M (2006) Hypoxia-inducible factor-1 (HIF-1). Mol Pharmacol 70(5):1469–1480 28. Li X, Lee Y, Kang Y, Dai B, Perez MR, Pratt M, Koay EJ, Kim M, Brekken RA, Fleming JB (2019) Hypoxia-induced autophagy of stellate cells inhibits expression and secretion of lumican into microenvironment of pancreatic ductal adenocarcinoma. Cell Death Differ 26:382–393 29. Gleissner CA, Galkina E, Nadler JL, Ley K (2007) Mechanisms by which diabetes increases cardiovascular disease. Drug Discov Today Dis Mech 4(3):131–140 30. Cerychova R, Pavlinkova G (2018) HIF-1, metabolism, and diabetes in the embryonic and adult heart. Front Endocrinol (Lausanne) 9:460 31. Bohuslavova R, Skvorova L, Sedmera D, Semenza GL, Pavlinkova G (2013) Increased susceptibility of HIF-1α heterozygous-null mice to cardiovascular malformations associated with maternal diabetes. J Mol Cell Cardiol 60:129–141 32. Sada K, Nishikawa T, Kukidome D, Yoshinaga T, Kajihara N, Sonoda K, Senokuchi T, Motoshima H, Matsumura T, Araki E (2016) Hyperglycemia induces cellular hypoxia through production of mitochondrial ROS followed by suppression of aquaporin-1. PLoS One 11(7): e0158619 33. McGinn O, Gupta VK, Dauer P, Arora N, Sharma N, Nomura A, Dudeja V, Saluja A, Banerjee S (2017) Inhibition of hypoxic response decreases stemness and reduces tumorigenic signaling due to impaired assembly of HIF1 transcription complex in pancreatic cancer. Sci Rep 7 (1):7872 34. Li S, Vallet S, Sacco A, Roccaro A, Lentzsch S, Podar K (2019) Targeting transcription factors in multiple myeloma: evolving therapeutic strategies. Expert Opin Investig Drugs 28 (5):445–462 35. Lee KE, Spata M, Bayne LJ, Buza EL, Durham AC, Allman D, Vonderheide RH, Simon MC (2016) Hif1a deletion reveals pro-neoplastic function of B cells in pancreatic neoplasia. Cancer Discov 6(3):256–269

Role of Heat Shock Protein 90 in Diabetes and Pancreatic Cancer Management

11

Pinninti Santosh Sushma, Saimila Momin, and Gowru Srivani

Abstract

Heat shock protein 90 (Hsp90) plays a key role in the activation of client proteins, which implicate Hsp90 in various biological processes, specifically coordinating regulatory mechanisms in order to control their activity. One of the key regulators responsible for the upregulation of Hsp90 is heat shock factor (HSF1), whose primary role is to bind heat shock elements (HSEs) with Hsp90 promoters. HSF1 functions by interacting with the transcriptional programming of Hsp90 and with integrate biological signals to regulate levels of Hsp90, especially during times of stress. Furthermore, not only are these Hsp90 protein chaperones upregulated but they can also be released from pancreatic beta cells during pro-inflammatory circumstances. Additionally, Hsp90 interferes with survival and metastatic pathways that are associated with pancreatic cancer (PC) progression. Future investigations on protein chaperons that are associated with Hsp90 may lead to the identification of biomarkers for diseases such as diabetes and PCs and potentially lead to therapeutic strategies in management of these chronic diseases. Keywords

Heat shock proteins · Diabetes · Pancreatic cancer

P. S. Sushma (*) Department of Biotechnology, Dr. NTR University of Health Sciences, Vijayawada, Andhra Pradesh, India e-mail: [email protected] S. Momin Department of Bioscience and Biotechnology, Banasthali University, Vanasthali, Rajasthan, India Department of Biology, Emory University, Atlanta, GA, USA G. Srivani Department of Bioscience and Biotechnology, Banasthali University, Vanasthali, Rajasthan, India # Springer Nature Singapore Pte Ltd. 2019 G. P. Nagaraju, A. BM Reddy (eds.), Exploring Pancreatic Metabolism and Malignancy, https://doi.org/10.1007/978-981-32-9393-9_11

183

184

P. S. Sushma et al.

Abbreviations 17-DMAG ATP CTA DBD FAK HIF-1a HSEs HSF Hsp90 HSR IGF-IR IL-6 NK cells NTA PBMC PC T1D VEGF

11.1

17-(Dimethylaminoethylamino)-17-demethoxygeldanamycin Adenosine triphosphate C-terminal transactivation DNA-binding domain Focal adhesion kinase Hypoxia-induced factor-1α Heat shock elements Heat shock factor Heat shock protein 90 Heat shock response Insulin-like growth factor-1 receptor Interleukin-6 Natural killer cells N-terminal transactivation Peripheral blood mononuclear cell Pancreatic cancer Type 1 diabetes Vascular endothelial growth factor

Introduction

Hsp90 (heat shock protein 90) constitutes nearly 2% of cellular proteins and 6% of cellular proteins among stressed cells [1–4]. Hsp90 levels usually depend on a master HSR (heat shock response) regulator and HSF1 (heat shock factor 1), subjected to regulatory processes. Hsp90, which is regulated by various mechanisms, influences transcription and is subjected to post-translational modification as well as modulation via co-chaperones. The human cells usually contain both Hsp90β and heat-inducible Hsp90α [5]; these molecules have an 86% amino acid sequence identity [6]. Despite the fact that these two molecules are highly evolutionarily conserved, these proteins exhibit a variety of different functions [7]. Surprisingly, Hsp90α may not even be necessary in mammals; however, on the other hand, Hsp90β participates in maintaining the viability. Hsp90α is involved in adaptive roles [8]. Hsp90 triggers the maturation of prominent signaling proteins like regulatory kinases [9–11], transcription factors, and hormone receptors [14]. Hsp90 is also associated with protein complexes [12] that suppress phenotypic variations [13–16]. Hsp90α and Hsp90β both interact with eukaryotic proteomes [17], which represents approximately 2000 proteins, among which around 725 interactions have been proven by protein to protein interaction in experimental results. This suggests that Hsp90 is highly necessary in a range of mechanisms in

11

Role of Heat Shock Protein 90 in Diabetes and Pancreatic Cancer Management

185

functional regulation. This review aims at examining the role of HSPs in diabetes and pancreatic cancer (PC).

11.2

Structure of Hsp90

Hsp90 contains a N-terminal dimerization domain which is primarily responsible for binding ATP to the middle domain through an unstructured charged linker. The N-terminal domains participate in the transient dimerization process by binding with ATP [18]. Then the C-terminal domain is responsible for protein dimerization. ATP binding promotes the lid segment moment in N-terminal domains over the bound ATP [19]. These lid movements expose the surface residues, which play a key role in the dimerization of the N-terminal domains within Hsp90. Activity of ATPase in Hsp90 is initiated when the middle domain catalytic loops relocate to the open active state [20]. These loops possess an arginine residue which interacts with γ -phosphate and therefore leads to the promotion of ATP hydrolysis by Hsp90. The catalytic loop confirmation may be modulated by binding with co-chaperones, such as Aha1, which activates the activity of ATPase [21]. The conformational changes that occur represent the rate-limiting step of the Hsp90 chaperone cycle. The molecular details regarding Hsp90 chaperone cycle activation and maturation are yet to be investigated.

11.3

Activation of HSF1 and the Heat Shock Response

HSF proteins regulate HSR [22–24], which can be induced by various stimuli like elevated temperature levels, viral and bacterial infections, and even oxidative stress [25]. Being a primary regulator of HSR and Hsp90 client protein, examining the functions of HSF1 is necessary in Hsp90 regulation. HSF proteins contain four essential components: DBD (DNA-binding domain), a trimerization domain, HR-C region, and CTA (C-terminal transactivation) [22, 23, 26–28]. Yeasts possess an NTA (N-terminal transactivation) domain [29]. The DNA-binding domain of these HSFs binds to HSEs (heat shock elements) that contain numerous nGAAn units [30]. The arrangement of these units leads to the promotion of cooperativity within the binding of HSF trimers. The trimerization domain in HSF constitutes of a couple of heptad repeats (HR-A and HR-B), forming a triple-stranded α-helical coil [31]. In Drosophila as well as mammals, HSF1 third repeat (HR-C) is known for intramolecular interactions to aid in maintaining HSF1 in monomeric form, which can be counteracted by stress [32– 34]. Trimerization is triggered by DBD intermolecular interactions, such as tryptophan and phenylalanine residues [35], and mammals contain two cysteine residues, which allows it to form disulfide bonds in response to stress [35, 36]. Elevated temperature levels result in hyperphosphorylation of particular proteins. This increases transcriptional programming when HSF trimer binds to the nGGAn units [37, 38]. When HSF1 trimer binds to HSEs that contain four phosphorylation sites at

186

P. S. Sushma et al.

Ser303, Ser307, as well as Ser363, HSF1 becomes repressive to transcription, but this process can be overwritten during stress conditions [39, 40], suggesting that this system is stress sensitive [26]. In yeast, an unstructured NTA domain [41] mediates stimulation of HSF1, suggesting its role as a negative regulator of the CTA domain; in other words, it is unable to activate stress-mediated HSF1 [42, 43]. The CTA domain encourages increased concentrations of promoter activity 15–33  C. On the other hand, transient activity, which is modulated by the NTA domain, is generally activated between 34.5 and 39  C [44]. During stress conditions, the levels of denaturation in proteins increases [28], triggering cytoplasmic non-DNA-binding HSF1 conversion to a homotrimer, which increases DNA-binding activity. Thus, HSF1 that is dispensed in association with Hsp90 [45–47] undergoes homotrimerization [34] as well as translocation to the nuclear region upon binding to HSE [48–50]. In the next phase, series of phosphorylation processes transform HSF1 trimers to activate transcription factors [26, 27], which results in upregulation of Hsp90 along with other chaperones including Hsp70, Hsp27, and Hsp40 [51].

11.3.1 HSPs and Diabetes Type 1 diabetes (T1D) is an autoimmune disease mediated by T-cell responses, in which pancreatic β cells, which produce insulin, are destroyed by the immune system. The shortage of identifiable biomarkers and disease-modifying therapies have become an obstacle in the management of T1D [52]. Studies have primarily focused on β cell stress pathways, like stress associated with endoplasmic reticulum as well as oxidative stress, which can potentially play a major role in further stimulating the immune system and accelerating T1D progression [53, 54]. The activation of such pathways precedes the advances in clinically detectable hyperglycemia [55], making the study of β cell health-based biomarkers and their application in diabetes an attractive option in both the prediction and therapeutic strategies of T1D. Islet Hsp90 levels found in experiments, conducted with NOD mice before hyperglycemia onset [56], were elevated. The expression of intracellular Hsp90 was also found to be upregulated, especially in response to environmental factors like reactive oxygen species, hypoxia, and irradiation [57]. Additionally, cadaveric human islets as well as pancreatic β cell lines discharged higher levels of Hsp90 during pro-inflammatory cytokine treatment [58, 59]. Pediatric populations with T1D exhibited higher serum levels of Hsp90 (alpha cytoplasmic form) in comparison with age- and gender-matched controls [56]. Furthermore, increased levels of circulating anti-Hsp90 autoantibodies were recognized in T1D patients [60]. Protein markers associated with ER stress were found to be increased in islets among individuals diagnosed with T1D [61]. ER stress promotes T1D development in NOD mice [62], whereas ER stress mitigation with chemical chaperones prevents T1D development in NOD mice [63]. In the pancreas, pro-inflammatory cytokines are released by macrophages, NK cells, as well as T cells during insulitis [59]. Ex vivo therapy of cadaveric islet cells as well as β cell lines with cytokines led to ER

11

Role of Heat Shock Protein 90 in Diabetes and Pancreatic Cancer Management

187

stress [64]. In contrast, previous studies have reported that pro-inflammatory cytokines released from β cells induce Hsp90 [58], and therefore the expression of the Hsp90 protein becomes increased among islets within prediabetic NOD mice when ER stress conditions arise. Investigations revealed blood Hsp90 levels also increased among pediatric patients during T1D diagnosis [56]. The elevated Hsp90 during diagnosis of T1D needs to be tested in order to establish whether it serves as a strong biomarker of T1D. Hsp90 is found to be expressed in several tissues, specifically immune cells. Studies have also demonstrated elevations in serum Hsp90 levels during both autoimmune and inflammatory conditions. In a cohort study, individuals diagnosed with bullous pemphigoid, along with an increase of peripheral blood mononuclear cells (PBMC), had elevated Hsp90 levels compared to the control group [65]. PBMC Hsp90 levels were also elevated in individuals diagnosed with systemic lupus erythematosus. Additionally, elevated PBMC Hsp90 levels are also strongly correlated with ascending levels of IgG autoantibodies to Hsp90 [66]. The elevations of serum Hsp90 levels promote a generalized inflammation condition within patients that have β cell autoimmunity. Studies were also conducted to investigate whether there is an association between Hsp90 levels with demographic and clinical variables. Future longitudinal analysis is needed to test this approach. Surprisingly, there is a negative correlation between Hsp90 levels and age in autoantibody-positive subjects. Additionally, there was also a negative trend with respect to age among progressors to T1D. These correlations are similar with previous studies that have investigated associations between serum concentrations and age. Hsp90 levels recorded in children, within the age group of 4–15 years, with recent onset of T1D, [56] served as an indication of aggressive B cell autoimmunity among this age group [67, 68]. A strong association between age and inflammation is well-established [69]. In a study, Hsp90 inhibitors exhibited therapeutic efficacy, playing the role as a senolytic agent in rodents to extend their lifespan [70]. The elevated serum Hsp90 levels in older individuals suggest that the elevated inflammation connected with aging, whereas the absence of age-associated increases is reporting Hsp90 as potential biomarker in detecting pathological inflammatory progressions among younger persons. The process of targeting oncogenic signaling pathways can lead to improvements in PC therapy. Hsp90 is recognized as a crucial component for the proper functioning of oncogenic kinases as well as signaling intermediates [71–73]. The transcription factors like HIF-1α and STAT3 are identified as client proteins of Hsp90 [73, 74]. Hsp90 is highly expressed among cancer cells [75]; studies indicate that Hsp90 inhibitors exhibit the highest affinity of binding to Hsp90 in malignant cells, in comparison with healthy cells [76]. The inhibition of Hsp90 may target signaling intermediates, including transcription factors, which can lead to reductions in PC growth and angiogenesis. To this date, geldanamycin and 17-(dimethylaminoethylamino)-17-demethoxygeldanamycin (17-DMAG) are considered to be the best inhibitors of Hsp90. These inhibitors are still being investigated but are involved in phase I/II trials of clinical studies [77–79]. However, current synthetic inhibitors that are being developed also have promising results as for cancer treatment [80, 81].

188

P. S. Sushma et al.

The Hsp90 inhibition interferes with IGF-I/IGF-IR signaling within cancer cells. During initial experiments, cells pretreated with 17-AAG inhibited IGF-I-mediated activation of both Erk1/2 and Akt signaling [82]. The STAT3 phosphorylation decreased IGF-I activation, suggesting STAT3 is not associated with the IGF-IR signaling cascade. IRS-1, which is upstream of Erk and Akt, is found to respond to 17-AAG during phosphorylation impairment. According to subsequent experiments, Akt and Erk1/2 inhibition by 17-AAG was dose-dependent, in which a dose of 100 nmol/L can lead to a reduction in phosphorylation substrates. These experiments strongly suggest that inhibiting Hsp90 impairs downstream signaling of the IGF-IR signaling pathway in PC. Inhibiting Hsp90 may also disrupt the IL-6/STAT3 pathway in PC [83, 84]. This pathway is activated by Hsp90 and is involved in metastasis of PC [85]. IL-6mediated STATs activation with treatment of 17-AAG significantly decreased STAT3 activation and phosphorylation of STAT5. IL-6 stimulation leads to Erk1/ 2 activation, which is regulated by Hsp90 in malignant cells. Similarly, PI3K/Akt inhibition via inhibitors such as LY290004 did not affect phosphorylation of STATs upon IL-6 exposure. Therefore, Hsp90 inhibition can directly disrupt IL-6- mediated activation in STAT3 and STAT5 in PC cells.

11.3.2 Effect of Hsp90 Inhibition on IL-6 and Hypoxia-Mediated Induction of HIF-1a HIF-1α is a transcription factor known to be a key promoter of cancer growth as well as angiogenesis in many cancers, including PC [82, 86, 87]. HIF-1α is necessary for complex formation with STAT3, in order to ultimately facilitate target gene expression such as VEGF [88, 89]. Hence, Hsp90 inhibition affects the function of HIF-1α either through direct inhibition or by inhibition of STAT3. Furthermore, studies showed that 17-AAG decreased the levels of nuclear HIF-1α protein in PC cells. HIF-1α induction by hypoxia was prevented by 17-AAG. Similarly, activation with IL-6 elevated nuclear HIF-1α content. The response decreased via Hsp90 inhibition with17-AAG. Hsp90 inhibition also decreased HIF-1α levels by decreasing its expression, which was experimentally identified by real-time PCR. Hsp90 inhibitor 17-AAG decreased VEGF-A expression and increased IGF-1 expression under hypoxic conditions, as well as IGF-I stimulation [82]. So Hsp90 inhibitors can be used in the reduction of VEGF-mediated angiogenesis in PC. Previous studies concluded that HIF-1α function may be inhibited by blocking Hsp90. Additionally, HIF-1α regulates cell motility in cancers [90]. Focal adhesion kinase (FAK), the chief mediator of cell motility, is a client protein of Hsp90. 17-AAG therapy can inhibit FAK phosphorylation in PC. Additionally, IGF-Imediated cancer cell motility was found to be impaired by 17-AAG. So Hsp90 inhibition can reduce PC metastasis by blocking invasive and migratory pathways.

11

Role of Heat Shock Protein 90 in Diabetes and Pancreatic Cancer Management

189

11.3.3 Effect of Hsp90 Blockade on Growth of PC A s.c. xenograft model of PC (HPAF-II) was used to evaluate the effects of Hsp90 inhibition on cancer growth. When treated with 17-DMAG, cancer growth was inhibited. Treatment with 17-DMAG upregulated Hsp70 in cancer samples. The changes in expression and in FAK phosphorylation were not detectable. The growth inhibitory as well as antiangiogenic properties of 17-DMAG were studied in an orthotopic pancreatic tumor model. After treatment with 17-DMAG for 21 days, data analyses showed a decrease in orthotopic cancer growth. Similarly, STAT3 activation was diminished in tumor cells treated with 17-DMAG. Analyses suggested that IGF-IR as well as STAT3 could be prominent targets of Hsp90 blocking in PC. Studies additionally recognized IL-6/STAT3/HIF-1, a novel autocrine activation loop in pancreatic tumor cells, can also cause disruptions by blocking Hsp90. Studies using orthotopic tumor models suggested that targeting Hsp90 with inhibitors such as 17-DMAG can reduce cancer growth as well as tumor vascularization. These inferences can lead to a multifactorial approach in using Hsp90 inhibitors in PC therapy. Overall, IGF-IR turned out to be a remarkable target in cancer therapy. Hsp90 inhibitors can directly affect IGF-IR, by decreasing receptor phosphorylation upon ligand stimulation. Inhibition leads to decreased expression of IGF-IR, when given in the appropriate time frame with the correct dose. Thus, Hsp90 elicits acute and chronic inhibitory effects on IGF-IR. Hsp90 inhibition also leads to downregulation of IGF-IRh mRNA followed by slight increase in IGF-IRh mRNA, thereby suggesting a potential response. Additionally, increase of this precursor also results in Hsp90 inhibitor-mediated damage of pro-convertase furin/PC5 [91]. The effect of downregulation of IGF-IR via Hsp90 inhibitors was described in a few other studies; however, its effect on phosphorylation was not discussed. Terry et al. [92] revealed geldanamycin-derivate 17-AAG decreased the IGF-IR expression in synovial carcinoma cell lines and concluded that cancer cells that overexpress IGF-IR are virtually nonreactive to Hsp90-inhibiting therapy. These findings suggest that Hsp90 inhibitors can serve as strong therapeutic options for PC, especially since IGF-IR is overexpressed and active in this cancer [82, 93, 94]. Additionally, Hsp90 inhibitors affect PC growth and angiogenesis through another mechanism besides IGF-IR. This mechanism involves IL-6, which is a cytokine that can increase angiogenesis by increasing the expression of VEGF-A [84, 95, 96]. Tang et al. [83] discussed IL-6-mediated VEGF upregulation in PC cell lines. Masui et al. [84] showed a strong correlation between IL-6 receptor expression and VEGF expression in human pancreatic cancer models. Furthermore, IL-6 signaling in malignant cells requires STAT3 activation [97, 98]. IL-6-mediated STAT3/5 activation may be inhibited in pancreatic tumor cells through 17-AAG. An IL-6 autocrine activation loop was identified because both a hypoxic environment and recombinant IL-6 led to the increase in levels of IL-6 mRNA in pancreatic malignant cells. This autocrine (IL-6/STAT3/ HIF-1α) loop contributes to PC. Thus,

190

P. S. Sushma et al.

targeting Hsp90 may disrupt the IGF-IR system as well as the IL-6/STAT3/HIF-1α autocrine activation loop within PC. Additionally, 17-AAG inhibits STAT5 phosphorylation [98, 99]. The growth inhibitory as well as antiangiogenic impact of Hsp90-targeted therapy was experimentally tested in a s.c. xenograft model and orthotopic cancer model in pancreatic tumor cells. Hsp90 inhibitors like 17-DMAG reduced growth as well as vascularization of PCs. The main regulators of pancreatic tumor cell growth and angiogenesis are IGF-IR and STAT3. Both IGF-IR and STAT3 can be successfully inhibited by Hsp90 inhibitors. These results are critical for clinical trial research. The efficacy of Hsp90-targeted treatment is examined by evaluating and measuring client proteins in FNAC biopsies from PC subjects [100]. The antineoplastic effect of Hsp90 inhibitors on cancers, other than pancreatic malignancies, is half of the maximum tolerated dose. Effectively targeting oncogenic pathways can be accomplished without high-dose Hsp90 inhibitors. Currently, Hsp90 inhibitors and their role in metastasis are still unclear. Price et al. [101] showed that geldanamycin, in a breast cancer model, promoted bone metastases. In PC, the Hsp90 inhibitors block pro-migratory molecules and weaken migration of pancreatic tumor cells [102]. A daily regimen of 17-DMAG therapy can be more efficacious. Mice that were given 17-DMAG exhibited decreased lymph node metastasis, thereby suggesting that metastatic features of pancreatic tumor cells can be functionally impaired in vivo. In conclusion, Hsp90 inhibition impairs functions of IGF-IR and disrupts the IL-6/HIF-1α/STAT3 autocrine activation loop within PC. These concepts are strongly connected with reduction in cancer growth as well as vascularization in orthotopic pancreatic tumor models. Hsp90 inhibitors therefore serve as valuable assets for molecular therapies in PC treatment.

References 1. Picard D (2002) Heat-shock protein 90, a chaperone for folding and regulation. Cell Mol Life Sci CMLS 59(10):1640–1648 2. Whitesell L, Lindquist SL (2005) HSP90 and the chaperoning of cancer. Nat Rev Cancer 5 (10):761 3. Taipale M, Jarosz DF, Lindquist S (2010) HSP90 at the hub of protein homeostasis: emerging mechanistic insights. Nat Rev Mol Cell Biol 11(7):515 4. Finka A, Goloubinoff P (2013) Proteomic data from human cell cultures refine mechanisms of chaperone-mediated protein homeostasis. Cell Stress Chaperones 18(5):591–605 5. Ammirante M, Rosati A, Gentilella A, Festa M, Petrella A, Marzullo L, Pascale M, Belisario M, Leone A, Turco M (2008) The activity of hsp90α promoter is regulated by NF-κB transcription factors. Oncogene 27(8):1175 6. Gupta RS (1995) Phylogenetic analysis of the 90 kD heat shock family of protein sequences and an examination of the relationship among animals, plants, and fungi species. Mol Biol Evol 12(6):1063–1073 7. Zuehlke AD, Beebe K, Neckers L, Prince T (2015) Regulation and function of the human HSP90AA1 gene. Gene 570(1):8–16

11

Role of Heat Shock Protein 90 in Diabetes and Pancreatic Cancer Management

191

8. Grad I, Cederroth CR, Walicki J, Grey C, Barluenga S, Winssinger N, De Massy B, Nef S, Picard D (2010) The molecular chaperone Hsp90α is required for meiotic progression of spermatocytes beyond pachytene in the mouse. PLoS One 5(12):e15770 9. Pearl LH (2005) Hsp90 and Cdc37–a chaperone cancer conspiracy. Curr Opin Genet Dev 15 (1):55–61 10. Hořejší Z, Takai H, Adelman CA, Collis SJ, Flynn H, Maslen S, Skehel JM, de Lange T, Boulton SJ (2010) CK2 phospho-dependent binding of R2TP complex to TEL2 is essential for mTOR and SMG1 stability. Mol Cell 39(6):839–850 11. Takai H, Xie Y, de Lange T, Pavletich NP (2010) Tel2 structure and function in the Hsp90dependent maturation of mTOR and ATR complexes. Genes Dev 24(18):2019–2030 12. Makhnevych T, Houry WA (2012) The role of Hsp90 in protein complex assembly. Biochim Biophy Acta (BBA)-Mol Cell Res 1823(3):674–682 13. Rutherford SL, Lindquist S (1998) Hsp90 as a capacitor for morphological evolution. Nature 396(6709):336 14. Lindquist S (2009) Protein folding sculpting evolutionary change. In: Cold Spring Harbor symposia on quantitative biology: 2009. Cold Spring Harbor Laboratory Press, Woodbury, pp 103–108 15. Yahara I (1999) The role of HSP90 in evolution. Genes Cells 4(7):375–379 16. Williams TA, Fares MA (2010) The effect of chaperonin buffering on protein evolution. Genome Biol Evol 2:609–619 17. Zhao R, Davey M, Hsu Y-C, Kaplanek P, Tong A, Parsons AB, Krogan N, Cagney G, Mai D, Greenblatt J (2005) Navigating the chaperone network: an integrative map of physical and genetic interactions mediated by the hsp90 chaperone. Cell 120(5):715–727 18. Prodromou C, Panaretou B, Chohan S, Siligardi G, O'Brien R, Ladbury JE, Roe SM, Piper PW, Pearl LH (2000) The ATPase cycle of Hsp90 drives a molecular ‘clamp’via transient dimerization of the N-terminal domains. EMBO J 19(16):4383–4392 19. Ali MM, Roe SM, Vaughan CK, Meyer P, Panaretou B, Piper PW, Prodromou C, Pearl LH (2006) Crystal structure of an Hsp90–nucleotide–p23/Sba1 closed chaperone complex. Nature 440(7087):1013 20. Meyer P, Prodromou C, Hu B, Vaughan C, Roe SM, Panaretou B, Piper PW, Pearl LH (2003) Structural and functional analysis of the middle segment of hsp90: implications for ATP hydrolysis and client protein and cochaperone interactions. Mol Cell 11(3):647–658 21. Meyer P, Prodromou C, Liao C, Hu B, Roe SM, Vaughan CK, Vlasic I, Panaretou B, Piper PW, Pearl LH (2004) Structural basis for recruitment of the ATPase activator Aha1 to the Hsp90 chaperone machinery. EMBO J 23(3):511–519 22. Sorger PK, Pelham HR (1988) Yeast heat shock factor is an essential DNA-binding protein that exhibits temperature-dependent phosphorylation. Cell 54(6):855–864 23. Pirkkala L, Nykanen P, Sistonen L (2001) Roles of the heat shock transcription factors in regulation of the heat shock response and beyond. FASEB J 15(7):1118–1131 24. Wiederrecht G, Seto D, Parker CS (1988) Isolation of the gene encoding the S. cerevisiae heat shock transcription factor. Cell 54(6):841–853 25. Sorger PK (1991) Heat shock factor and the heat shock response. Cell 65(3):363–366 26. Anckar J, Sistonen L (2011) Regulation of HSF1 function in the heat stress response: implications in aging and disease. Annu Rev Biochem 80:1089–1115 27. Holmberg CI, Tran SE, Eriksson JE, Sistonen L (2002) Multisite phosphorylation provides sophisticated regulation of transcription factors. Trends Biochem Sci 27(12):619–627 28. Voellmy R (2004) On mechanisms that control heat shock transcription factor activity in metazoan cells. Cell Stress Chaperones 9(2):122 29. Nieto-Sotelo J, Wiederrecht G, Okuda A, Parker CS (1990) The yeast heat shock transcription factor contains a transcriptional activation domain whose activity is repressed under nonshock conditions. Cell 62(4):807–817 30. Sakurai H, Enoki Y (2010) Novel aspects of heat shock factors: DNA recognition, chromatin modulation and gene expression. FEBS J 277(20):4140–4149

192

P. S. Sushma et al.

31. Peteranderl R, Rabenstein M, Shin Y-K, Liu CW, Wemmer DE, King DS, Nelson HC (1999) Biochemical and biophysical characterization of the trimerization domain from the heat shock transcription factor. Biochemistry 38(12):3559–3569 32. Farkas T, Kutskova YA, Zimarino V (1998) Intramolecular repression of mouse heat shock factor 1. Mol Cell Biol 18(2):906–918 33. Rabindran SK, Haroun RI, Clos J, Wisniewski J, Wu C (1993) Regulation of heat shock factor trimer formation: role of a conserved leucine zipper. Science 259(5092):230–234 34. Zuo J, Baler R, Dahl G, Voellmy R (1994) Activation of the DNA-binding ability of human heat shock transcription factor 1 may involve the transition from an intramolecular to an intermolecular triple-stranded coiled-coil structure. Mol Cell Biol 14(11):7557–7568 35. Lu M, Lee Y-J, Park S-M, Kang HS, Kang SW, Kim S, Park J-S (2009) Aromatic-participant interactions are essential for disulfide-bond-based trimerization in human heat shock transcription factor 1. Biochemistry 48(18):3795–3797 36. Ahn S-G, Thiele DJ (2003) Redox regulation of mammalian heat shock factor 1 is essential for Hsp gene activation and protection from stress. Genes Dev 17(4):516–528 37. Bulman AL, Nelson HCM (2005) Role of trehalose and heat in the structure of the C-terminal activation domain of the heat shock transcription factor. Proteins: Struct Funct Bioinf 58 (4):826–835 38. Pattaramanon N, Sangha N, Gafni A (2007) The carboxy-terminal domain of heat-shock factor 1 is largely unfolded but can be induced to collapse into a compact, partially structured state. Biochemistry 46(11):3405–3415 39. Kline MP, Morimoto RI (1997) Repression of the heat shock factor 1 transcriptional activation domain is modulated by constitutive phosphorylation. Mol Cell Biol 17(4):2107–2115 40. Knauf U, Newton EM, Kyriakis J, Kingston RE (1996) Repression of human heat shock factor 1 activity at control temperature by phosphorylation. Genes Dev 10(21):2782–2793 41. Cho HS, Liu CW, Damberger FF, Pelton JG, Nelson HCM, Wemmer DE (1996) Yeast heat shock transcription factor N-terminal activation domains are unstructured as probed by heteronuclear NMR spectroscopy. Protein Sci 5(2):262–269 42. Bonner JJ, Heyward S, Fackenthal DL (1992) Temperature-dependent regulation of a heterologous transcriptional activation domain fused to yeast heat shock transcription factor. Mol Cell Biol 12(3):1021–1030 43. Chen T, Li F, Chen B-S (2009) Cross-talks of sensory transcription networks in response to various environmental stresses. Interdiscip Sci: Comput Life Sci 1(2):162–162 44. Sorger PK (1990) Yeast heat shock factor contains separable transient and sustained response transcriptional activators. Cell 62(4):793–805 45. Ali A, Bharadwaj S, O'Carroll R, Ovsenek N (1998) HSP90 interacts with and regulates the activity of heat shock factor 1 in Xenopus oocytes. Mol Cell Biol 18(9):4949–4960 46. Bharadwaj S, Ali A, Ovsenek N (1999) Multiple components of the HSP90 chaperone complex function in regulation of heat shock factor 1 in vivo. Mol Cell Biol 19(12):8033–8041 47. Zou J, Guo Y, Guettouche T, Smith DF, Voellmy R (1998) Repression of heat shock transcription factor HSF1 activation by HSP90 (HSP90 complex) that forms a stress-sensitive complex with HSF1. Cell 94(4):471–480 48. Amin J, Ananthan J, Voellmy R (1988) Key features of heat shock regulatory elements. Mol Cell Biol 8(9):3761–3769 49. Pelham HR (1982) A regulatory upstream promoter element in the Drosophila hsp 70 heatshock gene. Cell 30(2):517–528 50. Xiao H, Lis JT (1988) Germline transformation used to define key features of heat-shock response elements. Science 239(4844):1139–1142 51. Heimberger T, Andrulis M, Riedel S, Stühmer T, Schraud H, Beilhack A, Bumm T, Bogen B, Einsele H, Bargou RC (2013) The heat shock transcription factor 1 as a potential new therapeutic target in multiple myeloma. Br J Haematol 160(4):465–476

11

Role of Heat Shock Protein 90 in Diabetes and Pancreatic Cancer Management

193

52. Watkins RA, Evans-Molina C, Blum JS, DiMeglio LA (2014) Established and emerging biomarkers for the prediction of type 1 diabetes: a systematic review. Transl Res 164 (2):110–121 53. Atkinson MA, Bluestone JA, Eisenbarth GS, Hebrok M, Herold KC, Accili D, Pietropaolo M, Arvan PR, Von Herrath M, Markel DS (2011) How does type 1 diabetes develop?: the notion of homicide or β-cell suicide revisited. Diabetes 60(5):1370–1379 54. Soleimanpour SA, Stoffers DA (2013) The pancreatic β cell and type 1 diabetes: innocent bystander or active participant? Trends Endocrinol Metab 24(7):324–331 55. Insel RA, Dunne JL, Atkinson MA, Chiang JL, Dabelea D, Gottlieb PA, Greenbaum CJ, Herold KC, Krischer JP, Lernmark Å (2015) Staging presymptomatic type 1 diabetes: a scientific statement of JDRF, the Endocrine Society, and the American Diabetes Association. Diabetes Care 38(10):1964–1974 56. Watkins RA, Evans-Molina C, Terrell JK, Day KH, Guindon L, Restrepo IA, Mirmira RG, Blum JS, DiMeglio LA (2016) Proinsulin and heat shock protein 90 as biomarkers of beta-cell stress in the early period after onset of type 1 diabetes. Transl Res 168:96-106. e101 57. Li W, Sahu D, Tsen F (2012) Secreted heat shock protein-90 (Hsp90) in wound healing and cancer. Biochim Biophys Acta (BBA)-Mol Cell Res 1823(3):730–741 58. Ocaña GJ, Pérez L, Guindon L, Deffit SN, Evans-Molina C, Thurmond DC, Blum JS (2017) Inflammatory stress of pancreatic beta cells drives release of extracellular heat-shock protein 90α. Immunology 151(2):198–210 59. Eizirik DL, Colli ML, Ortis F (2009) The role of inflammation in insulitis and β-cell loss in type 1 diabetes. Nat Rev Endocrinol 5(4):219 60. Qin H-Y, Mahon JL, Atkinson MA, Chaturvedi P, Lee-Chan E, Singh B (2003) Type 1 diabetes alters anti-hsp90 autoantibody isotype. J Autoimmun 20(3):237–245 61. Marhfour I, Lopez X, Lefkaditis D, Salmon I, Allagnat F, Richardson S, Morgan N, Eizirik DL (2012) Expression of endoplasmic reticulum stress markers in the islets of patients with type 1 diabetes. Diabetologia 55(9):2417–2420 62. Tersey SA, Nishiki Y, Templin AT, Cabrera SM, Stull ND, Colvin SC, Evans-Molina C, Rickus JL, Maier B, Mirmira RG (2012) Islet β-cell endoplasmic reticulum stress precedes the onset of type 1 diabetes in the nonobese diabetic mouse model. Diabetes 61(4):818–827 63. Engin F, Yermalovich A, Nguyen T, Hummasti S, Fu W, Eizirik DL, Mathis D, Hotamisligil GS (2013) Restoration of the unfolded protein response in pancreatic β cells protects mice against type 1 diabetes. Sci Transl Med 5(211):211ra156-211ra156 64. Brozzi F, Nardelli TR, Lopes M, Millard I, Barthson J, Igoillo-Esteve M, Grieco FA, Villate O, Oliveira JM, Casimir M (2015) Cytokines induce endoplasmic reticulum stress in human, rat and mouse beta cells via different mechanisms. Diabetologia 58(10):2307–2316 65. Tukaj S, Kleszczyński K, Vafia K, Groth S, Meyersburg D, Trzonkowski P, Ludwig RJ, Zillikens D, Schmidt E, Fischer TW (2013) Aberrant expression and secretion of heat shock protein 90 in patients with bullous pemphigoid. PLoS One 8(7):e70496 66. Ripley B, Isenberg D, Latchman D (2001) Elevated levels of the 90 kDa heat shock protein (hsp90) in SLE correlate with levels of IL-6 and autoantibodies to hsp90. J Autoimmun 17 (4):341–346 67. Greenbaum CJ, Beam CA, Boulware D, Gitelman SE, Gottlieb PA, Herold KC, Lachin JM, McGee P, Palmer JP, Pescovitz MD (2012) Fall in C-peptide during first 2 years from diagnosis: evidence of at least two distinct phases from composite type 1 diabetes TrialNet data. Diabetes 61(8):2066–2073 68. Steck AK, Vehik K, Bonifacio E, Lernmark A, Ziegler A-G, Hagopian WA, She J, Simell O, Akolkar B, Krischer J (2015) Predictors of progression from the appearance of islet autoantibodies to early childhood diabetes: The Environmental Determinants of Diabetes in the Young (TEDDY). Diabetes Care 38(5):808–813 69. Goldberg EL, Dixit VD (2015) Drivers of age-related inflammation and strategies for healthspan extension. Immunol Rev 265(1):63–74

194

P. S. Sushma et al.

70. Fuhrmann-Stroissnigg H, Ling YY, Zhao J, McGowan SJ, Zhu Y, Brooks RW, Grassi D, Gregg SQ, Stripay JL, Dorronsoro A (2017) Identification of HSP90 inhibitors as a novel class of senolytics. Nat Commun 8(1):422 71. Isaacs JS, Xu W, Neckers L (2003) Heat shock protein 90 as a molecular target for cancer therapeutics. Cancer Cell 3(3):213–217 72. Neckers L (2002) Hsp90 inhibitors as novel cancer chemotherapeutic agents. Trends Mol Med 8(4):S55–S61 73. Zhang H, Burrows F (2004) Targeting multiple signal transduction pathways through inhibition of Hsp90. J Mol Med 82(8):488–499 74. Mabjeesh NJ, Post DE, Willard MT, Kaur B, Van Meir EG, Simons JW, Zhong H (2002) Geldanamycin induces degradation of hypoxia-inducible factor 1α protein via the proteosome pathway in prostate cancer cells. Cancer Res 62(9):2478–2482 75. Gooljarsingh LT, Fernandes C, Yan K, Zhang H, Grooms M, Johanson K, Sinnamon RH, Kirkpatrick RB, Kerrigan J, Lewis T (2006) A biochemical rationale for the anticancer effects of Hsp90 inhibitors: slow, tight binding inhibition by geldanamycin and its analogues. Proc Natl Acad Sci 103(20):7625–7630 76. Kamal A, Thao L, Sensintaffar J, Zhang L, Boehm MF, Fritz LC, Burrows FJ (2003) A highaffinity conformation of Hsp90 confers tumour selectivity on Hsp90 inhibitors. Nature 425 (6956):407 77. Banerji U, Judson I, Workman P (2003) The clinical applications of heat shock protein inhibitors in cancer-present and future. Curr Cancer Drug Targets 3(5):385–390 78. Heath EI, Hillman DW, Vaishampayan U, Sheng S, Sarkar F, Harper F, Gaskins M, Pitot HC, Tan W, Ivy SP (2008) A phase II trial of 17-allylamino-17-demethoxygeldanamycin in patients with hormone-refractory metastatic prostate cancer. Clin Cancer Res 14 (23):7940–7946 79. Nowakowski GS, McCollum AK, Ames MM, Mandrekar SJ, Reid JM, Adjei AA, Toft DO, Safgren SL, Erlichman C (2006) A phase I trial of twice-weekly 17-allylamino-demethoxygeldanamycin in patients with advanced cancer. Clin Cancer Res 12(20):6087–6093 80. Biamonte MA, Shi J, Hong K, Hurst DC, Zhang L, Fan J, Busch DJ, Karjian PL, Maldonado AA, Sensintaffar JL (2006) Orally active purine-based inhibitors of the heat shock protein 90. J Med Chem 49(2):817–828 81. Zhang L, Fan J, Vu K, Hong K, Le Brazidec J-Y, Shi J, Biamonte M, Busch DJ, Lough RE, Grecko R (2006) 7 ‘-substituted benzothiazolothio-and pyridinothiazolothio-purines as potent heat shock protein 90 inhibitors. J Med Chem 49(17):5352–5362 82. Stoeltzing O, Liu W, Reinmuth N, Fan F, Parikh AA, Bucana CD, Evans DB, Semenza GL, Ellis LM (2003) Regulation of hypoxia-inducible factor-1α, vascular endothelial growth factor, and angiogenesis by an insulin-like growth factor-I receptor autocrine loop in human pancreatic cancer. Am J Pathol 163(3):1001–1011 83. Tang R-F, Wang S-X, Zhang F-R, Peng L, Xiao Y, Zhang M (2005) Interleukin-1alpha, 6 regulate the secretion of vascular endothelial growth factor A, C in pancreatic cancer. Hepatobiliary Pancreat Dis Int: HBPD INT 4(3):460–463 84. Masui T, Hosotani R, Doi R, Miyamoto Y, Tsuji S, Nakajima S, Kobayashi H, Koizumi M, Toyoda E, Tulachan SS (2002) Expression of IL-6 receptor in pancreatic cancer: involvement in VEGF induction. Anticancer Res 22(6C):4093–4100 85. Wei D, Le X, Zheng L, Wang L, Frey JA, Gao AC, Peng Z, Huang S, Xiong HQ, Abbruzzese JL (2003) Stat3 activation regulates the expression of vascular endothelial growth factor and human pancreatic cancer angiogenesis and metastasis. Oncogene 22(3):319 86. Büchler P, Reber HA, Büchler M, Shrinkante S, Büchler MW, Friess H, Semenza GL, Hines OJ (2003) Hypoxia-inducible factor 1 regulates vascular endothelial growth factor expression in human pancreatic cancer. Pancreas 26(1):56–64 87. Zhong H, De Marzo AM, Laughner E, Lim M, Hilton DA, Zagzag D, Buechler P, Isaacs WB, Semenza GL, Simons JW (1999) Overexpression of hypoxia-inducible factor 1α in common human cancers and their metastases. Cancer Res 59(22):5830–5835

11

Role of Heat Shock Protein 90 in Diabetes and Pancreatic Cancer Management

195

88. Gray MJ, Zhang J, Ellis LM, Semenza GL, Evans DB, Watowich SS, Gallick GE (2005) HIF-1α, STAT3, CBP/p300 and Ref-1/APE are components of a transcriptional complex that regulates Src-dependent hypoxia-induced expression of VEGF in pancreatic and prostate carcinomas. Oncogene 24(19):3110 89. Xu Q, Briggs J, Park S, Niu G, Kortylewski M, Zhang S, Gritsko T, Turkson J, Kay H, Semenza GL (2005) Targeting Stat3 blocks both HIF-1 and VEGF expression induced by multiple oncogenic growth signaling pathways. Oncogene 24(36):5552 90. Krishnamachary B, Berg-Dixon S, Kelly B, Agani F, Feldser D, Ferreira G, Iyer N, LaRusch J, Pak B, Taghavi P (2003) Regulation of colon carcinoma cell invasion by hypoxia-inducible factor 1. Cancer Res 63(5):1138–1143 91. Stawowy P, Kallisch H, Kilimnik A, Margeta C, Seidah NG, Chrétien M, Fleck E, Graf K (2004) Proprotein convertases regulate insulin-like growth factor 1-induced membrane-type 1 matrix metalloproteinase in VSMCs via endoproteolytic activation of the insulin-like growth factor-1 receptor. Biochem Biophys Res Commun 321(3):531–538 92. Terry J, Lubieniecka JM, Kwan W, Liu S, Nielsen TO (2005) Hsp90 inhibitor 17-allylamino17-demethoxygeldanamycin prevents synovial sarcoma proliferation via apoptosis in in vitro models. Clin Cancer Res 11(15):5631–5638 93. Nair PN, De Armond DT, Adamo ML, Strodel WE, Freeman JW (2001) Aberrant expression and activation of insulin-like growth factor-1 receptor (IGF-1R) are mediated by an induction of IGF-1R promoter activity and stabilization of IGF-1R mRNA and contributes to growth factor independence and increased survival of the pancreatic cancer cell line MIA PaCa-2. Oncogene 20(57):8203 94. Bergmann U, Funatomi H, Yokoyama M, Beger HG, Korc M (1995) Insulin-like growth factor I overexpression in human pancreatic cancer: evidence for autocrine and paracrine roles. Cancer Res 55(10):2007–2011 95. Chang C-Y, Li M-C, Liao S-L, Huang Y-L, Shen C-C, Pan H-C (2005) Prognostic and clinical implication of IL-6 expression in glioblastoma multiforme. J Clin Neurosci 12(8):930–933 96. Nilsson MB, Langley RR, Fidler IJ (2005) Interleukin-6, secreted by human ovarian carcinoma cells, is a potent proangiogenic cytokine. Cancer Res 65(23):10794–10800 97. Loeffler S, Fayard B, Weis J, Weissenberger J (2005) Interleukin-6 induces transcriptional activation of vascular endothelial growth factor (VEGF) in astrocytes in vivo and regulates VEGF promoter activity in glioblastoma cells via direct interaction between STAT3 and Sp1. Int J Cancer 115(2):202–213 98. Dudley AC, Thomas D, Best J, Jenkins A (2005) A VEGF/JAK2/STAT5 axis may partially mediate endothelial cell tolerance to hypoxia. Biochem J 390(2):427–436 99. Klampfer L (2006) Signal transducers and activators of transcription (STATs): novel targets of chemopreventive and chemotherapeutic drugs. Curr Cancer Drug Targets 6(2):107–121 100. Rubio-Viqueira B, Mezzadra H, Nielsen ME, Jimeno A, Zhang X, Iacobuzio-Donahue C, Maitra A, Hidalgo M, Altiok S (2007) Optimizing the development of targeted agents in pancreatic cancer: tumor fine-needle aspiration biopsy as a platform for novel prospective ex vivo drug sensitivity assays. Mol Cancer Ther 6(2):515–523 101. Price JT, Quinn JM, Sims NA, Vieusseux J, Waldeck K, Docherty SE, Myers D, Nakamura A, Waltham MC, Gillespie MT (2005) The heat shock protein 90 inhibitor, 17-allylamino-17demethoxygeldanamycin, enhances osteoclast formation and potentiates bone metastasis of a human breast cancer cell line. Cancer Res 65(11):4929–4938 102. Bagatell R, Beliakoff J, David CL, Marron MT, Whitesell L (2005) Hsp90 inhibitors deplete key anti-apoptotic proteins in pediatric solid tumor cells and demonstrate synergistic anticancer activity with cisplatin. Int J Cancer 113(2):179–188

Insulin Resistance Is a Common Core Tethered to Diabetes and Pancreatic Cancer Risk

12

Henu Kumar Verma and L. V. K. S. Bhaskar

Abstract

The association between type 2 diabetes mellitus (T2DM) and pancreatic cancer (PC) is complex bidirectional. Recent epidemiological data suggest that about 80% of PC patients have impaired glucose tolerance (IGT) or new onset of diabetes. Obesity, smoking, family history, genetic factors, insulin resistance (IR), and chronic pancreatitis are significant risk factors for PC. Several lines of evidences demonstrated that the production of insulin increases with increasing IR in diabetes patients. Hyperinsulinemia and IR are involved in the development of PC. Majority of studies suggested that risk ratio for diabetes in associated PC was higher in patients with small periods of diabetes as compared those with long periods of diabetes. Several studies also reported that antidiabetic drugs decrease risk for PC and increase survival in diabetes patients. In this chapter we describe the link between insulin resistance and pancreatic cancer risk. Keywords

Diabetes · Obesity · Insulin resistance · Biomarkers · Pancreatic cancer

Abbreviations ADA ADM AGE BMI

American Diabetes Association Adrenomedullin Advanced glycation end product Body mass index

H. K. Verma Stem Cell Laboratory, Institute of Endocrinology and Oncology, Naples, Italy L. V. K. S. Bhaskar (*) Guru Ghasidas Vishwavidyalaya, Bilaspur, India e-mail: [email protected] # Springer Nature Singapore Pte Ltd. 2019 G. P. Nagaraju, A. BM Reddy (eds.), Exploring Pancreatic Metabolism and Malignancy, https://doi.org/10.1007/978-981-32-9393-9_12

197

198

H. K. Verma and L. V. K. S. Bhaskar

CI CRP DPP-4 DZ FBG GCKR GI GLP-1 GWAS HbA1c IGF IGF-1 IGF-2 IGFBP-1 IGFBP-3 IGT IR ObR OR PC PCCC PDAC PPARG RAGE RR SHH sTNFR1 sTNF-R2 T2DM WHO

12.1

Confidence interval C-reactive protein Dipeptidyl peptidase-4 Thiazolidinedione Fasting blood glucose Glucokinase regulator gene Glucose intolerance Glucagon-like peptide-1 Genome-wide association studies Glycated hemoglobin Insulin-like growth factor Insulin-like growth factor 1 Insulin-like growth factor 2 Insulin-like growth factor-binding protein 1 Insulin-like growth factor-binding protein 3 Impaired glucose tolerance Insulin resistance Leptin receptor Odds ratio Pancreatic cancer Pancreatic Cancer Cohort Consortium Pancreatic ductal adenocarcinoma Peroxisome proliferator-activated receptor gamma Receptor for advanced glycation end products Risk ratio Sonic hedgehog Soluble tumor necrosis factor receptor 1 Soluble tumor necrosis factor receptor 2 Type 2 diabetes mellitus World Health Organization

Introduction

Pancreatic cancer (PC) is the fourth most common cause of cancer-related deaths in the United States [1]. PC is commonly diagnosed at advanced stages with a 5-year overall survival rate of only ~ 5–8% [2, 3]. It is expected that PC becomes the second most significant cause of cancer deaths by the year 2020 [4]. As there are no effective chemo- and radiotherapy treatment strategies and a silent feature of the disease, majority of the individuals diagnosed with PC die of this lethal disease [5]. Potential risk factors for PC include chronic pancreatitis, age (usually >40 years), gender (more prevalent in males), genetic factors, family history, smoking, excessive drinking, obesity, and diabetes [6–11]. In addition, epidemiological studies have associated type 2 diabetes mellitus (T2DM) with the PC risk [12, 13]. Some of the early symptoms of PC are nonspecific and overlap with the T2DM [14]. Quantitative

12

Insulin Resistance Is a Common Core Tethered to Diabetes and Pancreatic. . .

199

analysis of the timing of the genetic evolution of PC indicated a broad 20-year time window from a precursor lesion to develop distant metastases [15]. Previous studies suggested increased incidence of PC in diabetes patients. A retrospective case– control study suggested that about 68% of patients with PC had diabetes and 40% developed diabetes 3 years prior than the diagnosis of PC [16]. Like T2DM patients, patients with progressive PC express several metabolic abnormalities such as inhibition of glucose production, deregulation of hepatic glucose metabolism, and insulin resistance (IR). The progression of diabetes in patients with PC is expected as a secondary factor due to failure of pancreatic β-cell function and subsequent to the progression to peripheral IR [17]. Several epidemiological studies reported high incidence of PC with T2DM with a comparative risk ratio (RR) ranging from 1.6 to 2.0 [12, 18]. Hence in 74% of PC patients with diabetes, the diagnosis of diabetes ensued