molecular mechanisms of nanoliposomal c6-ceramide …

169
The Pennsylvania State University The Graduate School College of Medicine MOLECULAR MECHANISMS OF NANOLIPOSOMAL C6-CERAMIDE-INDUCED CELL DEATH IN CHRONIC LYMPHOCYTIC LEUKEMIA A Dissertation in Molecular Medicine by Ushma A. Doshi 2015 Ushma A. Doshi Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy December 2015

Upload: others

Post on 01-Feb-2022

3 views

Category:

Documents


0 download

TRANSCRIPT

The Pennsylvania State University

The Graduate School

College of Medicine

MOLECULAR MECHANISMS OF NANOLIPOSOMAL

C6-CERAMIDE-INDUCED CELL DEATH IN CHRONIC

LYMPHOCYTIC LEUKEMIA

A Dissertation in

Molecular Medicine

by

Ushma A. Doshi

2015 Ushma A. Doshi

Submitted in Partial Fulfillment of the Requirements

for the Degree of

Doctor of Philosophy

December 2015

ii

The dissertation of Ushma A. Doshi was reviewed and approved* by the following: Charles Lang Director of the Graduate Program Professor of Cellular and Molecular Physiology Dissertation Co-Adviser Co-Chair of Committee

Mark Kester Professor of Pharmacology, University of Virginia School of Medicine Dissertation Co-Adviser Co-Chair of Committee

Thomas P. Loughran Professor of Medicine, University of Virginia Cancer Center Hong-Gang Wang Lois High Berstler Professor, Pediatrics and Pharmacology Jin-Ming Yang Professor of Pharmacology

David Claxton Special Member Professor of Medicine

Richard R. Young Professor of Supply Chain Management; Business Administration

*Signatures are on file in the Graduate School.

iii

ABSTRACT

Chronic lymphocytic leukemia (CLL) is the most prevalent form of adult

leukemia in Western countries. Despite a high incidence, its pathogenesis is still

poorly understood, hence limiting treatment strategies. Furthermore, since CLL is

predominantly a disease of the elderly, numerous therapeutic strategies are

unsuitable due to limited physical fitness of the patient. Therefore, the CLL

remains incurable for most patients. Further research is needed to develop novel

therapeutic strategies.

Ceramide is a ‘tumor suppressor’ sphingolipid known to regulate

differentiation, senescence and cell cycle arrest. While a large body of work

reveals the mechanism of nanoliposomal ceramide (CNL)-induced cell death in

several types of cancers, the effect in CLL remains unclear. This study

investigates the effect of CNL in CLL and deciphers the key signaling

mechanisms mediating CNL-induced cell death. We have shown that CNL

selectively induces cell death in CLL cells by targeting the Warburg effect

through reducing levels of glyceraldehyde-3-phosphate dehydrogenase

(GAPDH), with no detrimental effects on normal peripheral blood mononuclear

cells. Additionally, CNL treatment results in tumor regression in an in vivo murine

model of CLL. Several reports in the literature have shown that signal transducer

and activator of transcription 3 (STAT3) is constitutively phosphorylated on

serine-727 in CLL and that STAT3 might be a therapeutic target in this disease.

We demonstrate that CNL suppresses STAT3 phosphorylation at both tyrosine-

705 and serine-727 by inhibiting multiple upstream kinases that include Bruton’s

tyrosine kinase, mitogen-activated protein kinase kinase and protein kinase C.

iv

This suppression in STAT3 phosphorylation and the subsequent downregulation

of STAT3 transcriptional activity mediates CNL-induced cell death in CLL. Recent

work in the literature has uncovered that STAT3 phosphorylated at serine-727

associates with mitochondrial components and regulates the respiratory

chain. Overactivation of mitochondrial STAT3 phosphorylated at serine-727

confers viability and stress protection to CLL cells. Our initial results demonstrate

that CNL treatment reduces mitochondrial STAT3 levels, which might also be

critical to the cell death induction.

Taken together, our results suggest that inhibition of glycolytic respiration

and inhibition of STAT3 transcriptional activity are key signaling mechanisms of

CNL-induced cell death in CLL cells. Additionally, we also speculate that

inhibition of STAT3-dependent mitochondrial respiration is also critical for

induction of cell death by CNL treatment. We conclude that CNL could potentially

be an effective therapy for CLL. Overall, this work emphasizes targeting the

sphingolipid pathway and development of sphingolipids-based therapeutics for

cancer.

v

TABLE OF CONTENTS

LIST OF FIGURES ........................................................................................................ vii

ABBREVIATIONS .......................................................................................................... ix

PREFACE ...................................................................................................................... xi

ACKNOWLEDGEMENTS .............................................................................................. xii

CHAPTER 1: Literature Review ................................................................................... 1

Sphingolipids .............................................................................................................. 1

Sphingolipid metabolism .......................................................................................... 1

Functions of Sphingolipids ....................................................................................... 5

Ceramide for cancer therapeutics ............................................................................... 8

Ceramide and induction of cell death ......................................................................11

Chemotherapy-induced ceramide generation by the de novo pathway ...................16

Chemotherapy-induced ceramide generation by sphingomyelinase pathway .........20

Nanotechnology-based drug delivery of ceramide ..................................................23

Chronic Lymphocytic Leukemia (CLL) ........................................................................30

Conclusions ...............................................................................................................34

CHAPTER 2: Nanoliposomal C6-ceramide target the Warburg effect in chronic

lymphocytic leukemia .................................................................................................35

Abstract .....................................................................................................................36

Introduction ................................................................................................................37

Materials and Methods ...............................................................................................40

Results .......................................................................................................................48

Discussion .................................................................................................................63

CHAPTER 3: STAT3 mediates nanoliposomal C6-ceramide-induced cell death in

chronic lymphocytic leukemia ....................................................................................69

Introduction ................................................................................................................69

Materials and Methods ...............................................................................................72

Results .......................................................................................................................78

Discussion ............................................................................................................... 103

vi

CHAPTER 4: Effect of nanoliposomal C6-ceramide on mitochondrial bioenergetics

and mitochondrial STAT3 in chronic lymphocytic leukemia .................................. 109

Introduction .............................................................................................................. 109

Methods and Materials ............................................................................................. 112

Results ..................................................................................................................... 116

Discussion ............................................................................................................... 123

CHAPTER 5 – Conclusions, Future directions and Therapeutic Potential ............ 127

Conclusions ............................................................................................................. 127

Use of nanoliposomal C6-ceramide as a therapy for CLL ..................................... 127

CNL targets the Warburg effect in CLL ................................................................. 129

STAT3 mediates CNL-induced cell death in CLL .................................................. 130

Effect of CNL on mitochondrial STAT3 ................................................................. 131

Effect of CNL on cellular bioenergetics in CLL ...................................................... 132

Future Directions...................................................................................................... 136

Summary and therapeutic potential .......................................................................... 138

References .................................................................................................................. 139

Appendix: Letters of Permission .................................................................................. 153

vii

LIST OF FIGURES

Figure 1-1 Fundamental structures of sphingolipids………...……………………...1

Figure 1-2 Sphingolipid metabolism pathways………………………………………4

Figure 1-3 Compartmentalization of sphingolipid metabolism…………………..…4

Figure 1-4 Strategies to alter the sphingolipid balance in cancer cells to

potentiate cytotoxicity of chemotherapeutic drugs…………………………………10

Figure 1-5 Points of intervention in the sphingolipid pathway……………………29

Figure 2-1 Nanoliposomal C6-ceramide selectively induces cell death in CLL

cells…..………………………………………………………………………………….49

Figure 2-2 Cell death induced by nanoliposomal C6-ceramide occurs through

caspase 3/7-independent necrosis…………………………………………………..52

Figure 2-3 C6-ceramide nanoliposomes target GAPDH in CLL…………………55

Figure 2-4 C6-ceramide targets the glycolytic pathway…………………………..59

Figure 2-5 Pharmacological and molecular confirmation that nanoliposomal C6-

ceramide targets the glycolytic pathway at the level of GAPDH………………….60

Figure 2-6 Nanoliposomal C6-ceramide displays anti-leukemic effect in a CLL

animal model………………………………………………………...…………………62

Figure 3-1A STAT3 is overexpressed in CLL cell lines and patient cells……….79

Figure 3-1B Knockdown of STAT3 induces cell death in CLL cells……………..80

Figure 3-1C STAT3 inhibition reduces viability of CLL cels………………………81

Figure 3-2A CNL suppresses the phosphorylation of STAT3 in CLL cell lines...83

Figure 3-2B CNL suppresses phosphorylation of STAT3 in CLL patient cells…84

Figure 3-2C CNL does affect cellular viability & STAT3 phosphorylation in

HEK293 cells…………………………………………………………………………...85

Figure 3-3ACNL-induced suppression of phosphorylation is specific to STAT3.86

Figure 3-3B Suppression of STAT3 phosphorylation is specifically an effect of

CNL and not other sphingolipids……………………………………………………..87

Figure 3-4A CNL induces necrotic cell death in CLL cell lines……………..……89

viii

Figure 3-4B CNL induces necrotic cell death in CLL patient cells……………….89

Figure 3-4C CNL-induced suppression of p-STAT3 precedes induction of cell

death…………………………………………………………………………………….90

Figure 3-4D CNL-induces early time point suppression of p-STAT3 in CLL

patient cells……………………………………………………………………………..91

Figure 3-5A CNL does not activate phosphatases……………………………......92

Figure 3-5B CNL suppresses the activity of BTK………………………………….94

Figure 3-5C BTK inhibitors suppress phosphorylation of STAT3………………..94

Figure 3-5D CNL suppresses the activity of MEK1/2 kinase…………………….95

Figure 3-5E MEK1/2 inhibitors suppress phosphorylation of STAT3……………96

Figure 3-5F CNL suppresses the activity of PKC………………………………….97

Figure 3-5G PKC inhibitor suppress phosphorylation of STAT3…………………97

Figure 3-6A CNL reduces the levels of STAT3-regulated genes………………..99

Figure 3-6B Reduction of STAT3 phosphorylation precedes reduction of Mcl-1

levels following CNL treatment……………………………………………………….99

Figure 3-6C CNL inhibits expression of luciferase in a STAT3 luciferase reporter

assay…………………………………………………………………………………..100

Figure 3-7 STAT3-C expressing cells are resistant to CNL-induced death…...102

Figure 4-1 CNL treatment results in accumulation of ceramide in the

mitochondria, decreased mitochondrial membrane potential, and generation of

ROS……………………………………………………………………………………118

Figure 4-2 CNL treatment results in cytosolic release of AIF…………………...120

Figure 4-3 CNL treatment dimishes levels of total STAT3 and p-STAT3 in the

mitochondria…………………………………………………………………………..121

Figure 4-4 CNL treatment suppresses the glycolytic flux of JVM-3 cells……...122

Figure 5-1 Molecular mechanisms of CNL-induced cell death in CLL cells…..136

ix

ABBREVIATIONS

AIF Apoptosis inducing factor

ATP Adenosine triphosphate

BCR B-cell receptor

BTK Bruton’s Tyrosine Kinase

CDase Ceramidase

CerS Ceramide synthase

CLL Chronic lymphocytic leukemia

CNL Nanoliposomal C6-ceramide

C6-ceramide D-erythro-hexanoyl-sphingosine

ECAR Extracellular acidification rate

ER Endoplasmic reticulum

ETC Electron transport chain

FB1 Fumonisin B1

GAPDH Glyceraldehyde 3-phosphate dehydrogenase

GCS Glucosylceramide synthase

GlcCer Glucosylceramide

GLUT1 Glucose transporter 1

HIF-1α Hypoxia inducible factor 1 alpha

LGL Large granular lymphocytic leukemia

MEK Mitogen-activated protein kinase kinase

MOMP Mitochondrial outer membrane permeabilization

OA Okadaic acid

OCR Oxygen consumption rate

PKC Protein kinase C

x

PV Pervanadate

ROS Reactive oxygen species

SMase Sphingomyelinase

SMS Sphingomyelin synthase

Sph Sphingosine

SphK Sphingosine kinase

SPT Serine palmitoyl transferase

STAT3 Signal Transducer and Activator of Transcription 3

S1P Sphingosine 1-phosphate

TNF Tumor necrosis factor

TNFR1 Tumor necrosis factor receptor 1

TRAIL TNF-related apoptosis-inducing ligand

TRAILR1 TRAIL receptor 1

TRAILR2 TRAIL receptor 2

UGCG UDP-glucose ceramide glucosyltransferase

xi

PREFACE

I would like to recognize that Chapter 2 of my dissertation titled “C6-ceramide

nanoliposomses target the Warburg effect in chronic lymphocytic leukemia” is

derived from published data (PLoS One, 2013 Dec 19;8(12):e84648) for which I

am co-first author. This project was underway when I joined the Kester lab in

2011. Dr. Lindsay Ryland and I contributed towards the design of the

experiments within this particular chapter. Specifically, I was a major contributor

towards the following figures: Fig 2-1A, Fig. 2-3A, Fig. 2-4A and Fig 2-5. I am in

no way attempting to claim intellectual property over the design of all the

experiments within this particular chapter. Therefore, I would like to acknowledge

Dr. Lindsay Ryland and the other contributing authors for the work presented in

this chapter.

In addition, Dr. Lindsay Ryland performed experiments for Fig. 4-1 in

Chapter 4 of my dissertation titled “Effect of C6-ceramide Nanoliposomes on

mitochondrial bioenergetics and mitochondrial STAT3 in Chronic Lymphocytic

Leukemia”.

The remainder of my dissertation includes part of a published book

chapter for which I am first author (Chapter 1), manuscript for which I am first

author (Chapter 3), and manuscript in preparation for which I am first author

(Chapter 4).

xii

ACKNOWLEDGEMENTS

I would like to thank, first and foremost, my mentor and advisor Dr. Mark

Kester for giving me an opportunity to work in his laboratory and for guiding me

through this memorable journey. I owe my achievements to his incredible

support, guidance and his faith in my work. I would like to express my gratitude

towards him for encouraging me to pursue the MBA degree during the last two

years of my thesis work. Thank you for your patience and flexibility with my

unnatural working hours. I would also like to thank Dr. Thomas Loughran for his

mentorship throughout my graduate years. I would like to specifically thank Dr.

Charles Lang for his immense support and belief in my abilities as a graduate

student. I would not have been a part of Penn State College of Medicine, if it

were not for your conviction in my potential as a graduate student. I would like to

extend special thanks to Dr. HG Wang and Dr. Claxton for always being so

approachable and providing insightful comments on my work. In addition, I would

like to thank Dr. Jin-Ming Yang and Dr. Robert Young for being very supportive

committee members. All my professors at Penn State College of Medicine,

especially Dr. Kent Vrana, Dr. Ralph Keil and Dr. Xin Liu have extended their

support and guidance in my journey. I am indebted to past and present Kester

lab members for making the lab my second home. Special thanks to Dr. Todd

Fox and Dr. Su-Fern Tan for their valuable guidance in my work and their very

special friendship. Special thanks to Dr. Jeremy Haakenson, Dr. Jody Hankins

and Dr. Megan Young for their consistent support. I would like to thank Sriram

Shanmugavelandy for being a superb friend in lab. Thanks to Samuel Linton,

xiii

Tony Brown, Dr. Brian Barth and all other members in the Kester lab for this

spectacular journey. Special thanks to Taryn Dick, members of the Wang lab,

administrative staff of the Graduate Office at College of Medicine and the Penn

State Hershey Security.

I cannot imagine this journey without my friends and family. Vijay Kale,

Rameshwari Kale, Manmeet Raval and Darshan Trivedi – Hershey was indeed

the sweetest place on Earth for me because of your unfailing love and support.

Lastly, and most importantly, I am blessed with the best family. My best friend

and husband, Varun Prabhu, has been my guiding light and the best companion

in the last ten years. Thank you Varun, for your undying motivation, strength and

love. Special thanks to my sister, Kshama Doshi for being the source of my

strength, my best friend and my energy throughout. I heartily thank my parents

and my in laws for their love, constant support, motivation and their

understanding throughout my PhD years. I dedicate this work to my family.

Lastly, I would like to thank God for giving me the strength to dream big and

the grit to achieve them.

1

CHAPTER 1: Literature Review

Sphingolipids

Sphingolipids, a group of bioactive lipids, represent one of the eight major

classes of lipids [1]. This class of lipids are structurally characterized by a

sphingoid base backbone comprising of sphingosine most frequently, and the

presence of an amide-linked fatty acid and/or a headgroup attached to the

hydroxyl on carbon 1 (Fig. 1-1) [1].

Sphingolipid metabolism

Sphingolipid metabolism is a complex, compartmentalized and a highly

inter-connected system comprising of enzymes catalyzing the formation of

different classes of sphilgolipids. Ceramide is considered to be the central hub of

sphingolipid metabolism. The generic ‘ceramide’ is a family of more than 50

distinct molecular species with a base structure consisting of an acyl chain of

variable length attached to the sphingosine backbone [2]. Fig. 1-2 and Fig. 1-3

represent the sphingolipid metabolism pathway and the corresponding cellular

compartments.

Saturated bond forms

dihydroceramides

Figure 1-1 Fundamental structures of sphingolipids: Modified from Merrill et al. [1]. Sphingolipids are defined by having a sphingoid base (shown for sphingosine) that is often derivatized with an amide-linked fatty acid and/or headgroup of the general types shown.

2

Ceramide can be synthesized by an anabolic and a catabolic pathway. De

novo synthesis starts by serine palmitoyl transferase (SPT)-catalyzed

condensation of palmitate and serine to form 3-keto-dihydrosphingosine, which is

subsequently reduced to dihydrosphingosine followed by acylation by ceramide

synthases (CerS). Desaturases catalyze the formation of ceramide from

dihydroceramide. The endoplasmic reticulum (ER) and ER-associated

membranes are the site of de novo synthesis of ceramide [2].

The catabolic pathway of ceramide synthesis involves the conversion of

sphingomyelin to form ceramide by the action of acid sphingomyelinase (SMase)

residing in the outer membrane leaflet or neutral SMase in the inner leaflet of the

bilayer. Sphingomyelin transported to the lysosomes can also be converted to

ceramide by the action of lysosomal SMase. Ceramide can reversibly be

converted to sphingomyelin by the action of sphingomyelin synthase (SMS) in

the Golgi apparatus. The sphingomyelin produced in the Golgi can be

transported to the plasma membrane by vesicular transport where it can be

converted back to ceramide by the action of SMase as described earlier [2].

Ceramide can also be metabolized to glucosylceramide (GlcCer) in the

Golgi apparatus by the action of GlcCer synthase (GCS). GlcCer serves as the

precursor of complex glycosphingolipids in the Golgi. Reversibly, ceramide can

also be synthesized from GlcCer by the action of glucosyl ceramidase localized

in the lysosomal compartment [2]. Ceramide is also metabolized to ceramide-1-

phosphate by the action of ceramide kinase.

3

Another critical metabolism pathway is deacylation of ceramide to

sphingosine (Sph) by the action of ceramidases (CDase). There are five

ceramidases that are products of different genes: neutral CDase, acid CDase

and three forms of alkaline CDase [3]. These enzymes lie at a crucial juncture in

the sphingolipid pathway since, in conjunction with sphingosine kinases (SphK),

they balance the ceramide/sphingosine-1-phosphate (S1P) rheostat in cells.

Lysosomal acid CDase hydrolyses ceramide to form sphingosine (Sph), which is

favorably partitioned into the lysosomes due to its positive charge [2].

Furthermore, conversion of sphingosine to ceramide is mediated by CerS, which

forms a part of the de novo synthesis pathway of ceramide.

Sphingosine kinases 1 and 2 (SphK1 and SphK2) are the two sphingosine

kinase isozymes that catalyze the formation of S1P from sphingosine. It has

been postulated that SphK1 is present just outside the lysosomes ensuring

effective trapping of the Sph within the lysosomes by SphK1-mediated

phosphorylation. The presence of S1P phosphatases in the ER generates Sph

which can move among cellular biomembranes or which is eventually recycled to

form ceramide [2]. S1P lyase, which metabolizes S1P to ethanolamine

phosphate and hexadecenal, is the exit pathway in sphingolipid metabolism.

4

Figure 1-2 Sphingolipid metabolism pathways: Taken from Hannun et al. [2]

Figure 1-3 Compartmentalization of sphingolipid metabolism: Taken from Hannun et al. [2].

5

Functions of Sphingolipids

Sphingolipids function both as structural components of the cell and

mediators of cell signaling. As structural components of cellular biomembranes,

they play a critical role in regulating membrane fluidity and subdomain structure

of the lipid bilayer, especially lipid rafts [4]. Sphingolipids create lateral

differentiation of cellular membranes into a mosaic of structural domains with

unique molecular compositions. Lateral lipid assemblies are formed as a result of

varying miscibility of cell membrane-forming lipids like sphingolipids,

glycerolipids, and sterols. The protein content of such microdomains or rafts

characterize their function and serve as platforms for cellular events like signal

transduction, cell adhesion and protein sorting [5]. Sphingolipids like ceramides

affect the composition and properties of phospholipid bilayers by increasing the

order of the acyl chain in the bilayer, creating phase separation of ceramide-rich

and ceramide-poor domains and facilitating transition from bilayer to non-bilayer

structure [6]. In addition to cellular biomembranes, enzymatically-synthesized

ceramides also alter the properties of lipidic vesicles by inducing destabilization

of lipid bilayers through permeabilization and fusion [6]. Various biological

consequences follow ceramide-induced biophysical changes in cellular

biomembranes. For instance, the change in membrane fluidity after ceramide-

induced raft formation may result in modification of enzymatic activity of

membrane proteins, or may change the protein affinity for the membrane.

Ceramides also bind to specific sites in the target protein and alter enzymatic

activity. These target proteins are both membrane-bound proteins and other

6

cytoplasmic proteins transiently recruited to the bilayer where ceramides are

located [6].

In addition to structural roles, sphingolipids play a critical role in cellular

signaling. Extensive work has been done to elucidate the role of sphingolipids in

modulating cellular signaling. This section will broadly discuss the mechanisms

through which sphingolipids regulate cellular signaling mechanisms. Firstly,

sphingolipids act as ligands to receptors, initiating signaling pathways involved in

cellular processes like growth, adhesion, differentiation and migration. The

ligand-receptor interaction can be triggered in the same cell secreting the

sphingolipids or in neighboring cell types [7]. For instance, S1P is a ligand for a

family of G-protein coupled receptors called S1P receptors that regulate

biological processes such as cell proliferation, angiogenesis, migration, immune

cell trafficking and mitogenesis. Secondly, sphingolipids influence the properties

of receptors via specific lipid-protein interactions and regulating responsiveness

to external stimuli. For instance, it has been shown that the ligand-binding

capacity of human serotonin 1A receptors is impaired under glycosphingolipid-

depleted condition [8]. The authors speculate that this effect is due to a reduction

in the specific interaction of serotonin 1A receptors with membrane

glycosphingolipids [8]. Thirdly, as discussed earlier, sphingolipids alter the

biophysical properties of cellular biomembranes affecting the assembly of

membrane receptors and effector molecules in specific domains called rafts. This

sphingolipid-induced alteration in membrane biophysics regulates cellular

signaling [7]. Lipid rafts create a micro-environment suitable for effective receptor

7

activation by concentrating relevant kinases and protecting proteins from

phosphatases and other molecules that may diminish the signaling processes.

For example, lipid rafts are critical for immunoglobulin E signaling and T-cell

antigen receptor-mediated signaling [9]. Lastly, sphingolipids like ceramides and

S1P can act as direct signaling molecules in processes like proliferation,

differentiation, senescence, cell–cell interaction and transformation [7]. For

instance, intranuclear S1P has been shown to inhibit the enzymatic activities of

histone deacetylases, preventing the removal of acetyl groups from lysine

residues within histone tails. This work has shown the role of intranuclear S1P in

regulating gene expression [10].

Because sphingolipids are bioactive molecules and mediate several

cellular processes like proliferation, differentiation, migration, death and cell-cell

interaction, they have been implicated in the pathogenesis of several human

disorders. There is an abundance of literature reporting the role for sphingolipids

in inflammatory and immune responses, vascular function, neurodegeneration,

insulin signaling and diabetes, microbial pathogenesis and cancer pathogenesis

and therapy [11]. This dissertation is focused on studying the role of short-chain

ceramides as a potential therapy for chronic lymphocytic leukemia (CLL).

Furthermore, this dissertation also focuses on elucidating the key molecular

mechanisms which mediate the effect of short-chain ceramides in CLL cells.

8

Ceramide for cancer therapeutics

Sphingolipids have been implicated in cancer pathogenesis since they

mediate several cellular processes like proliferation, differentiation, migration,

death and cell-cell interaction. Numerous studies in the literature have reported a

dysregulation in sphingolipid metabolism in different types of cancers. The

dysregulated sphingolipid profile in cancer cells contributes to cancer progression

and metastasis, making it an ideal candidate for developing targeted

therapeutics. Moreover, sphingolipids may also serve as vital biomarkers for

cancer progression, as well as guide therapeutic regimens [12].

The role of sphingolipids in cancer pathogenesis and treatment has been

of particular interest to the sphingolipid research community. First reports

describing the involvement of sphingolipids in mediating apoptosis in cancer cells

came in early 1990s. It was reported that synthetic short chain ceramide analogs

like C2-ceramide induced cell death in HL60 leukemic cells and caused

internucleosomal DNA fragmentation [13]. Tumor necrosis factor α (TNFα) and

ionizing radiation induced apoptotic cell death in cancer cells, which was

mediated by sphingomyelin hydrolysis and subsequent ceramide generation [14,

15]. Apoptotic cell death through CD95 crosslinking in U937 cells also utilized the

sphingomyelin pathway and depended on ceramide production [16]. The first

body of work demonstrating the role of sphingolipids in mediating chemotherapy-

induced apoptosis in cancer cells was published in 1995, wherein the authors

showed that daunorubicin, a chemotherapeutic drug, induced apoptosis in P388

and U937 leukemia cells by elevating intracellular ceramide levels. This increase

9

in the intracellular ceramide pool was not due to the action of SMase, but rather

via activation of CerS, which increased de novo ceramide synthesis within cells

[17]. This revelation of a potential role of sphingolipid metabolism in mediating

chemotherapy-induced cytotoxicity generated immense interest in the scientific

community to decipher the connection between chemotherapy and sphingolipid

metabolism. Since then, a large body of research has been conducted to

establish an in-depth understanding of how sphingolipid metabolism mediates,

enhances, or impedes chemotherapy-induced cytotoxicity, with the goal of

identifying critical therapeutic targets and better therapeutic regimens for

management and cure of cancer.

Among the major sphingolipids that play a role in regulating cancer cell

fate, ceramide is termed as a “tumor suppressor lipid” because of its ability to

potentiate signaling cascades that lead to cell death. By contrast, sphingosine-1-

phosphate (S1P) is considered as a pro-survival lipid. Thus, in the context of

sphingolipids, the ceramide-S1P rheostat dictates cancer cell fate. Efforts

directed at altering the sphingolipid balance in cancer cells to induce cell death or

to potentiate cytotoxicity of chemotherapeutic drugs would aim at either elevating

pro-apoptotic sphingolipids, especially ceramide, or down-regulating pro-survival

sphingolipids such as sphingosine-1-phosphate. This can be achieved by: (i)

chemotherapy-induced synthesis of pro-apoptotic ceramides or breakdown of

pro-survival sphingolipids; (ii) disruption of ceramide metabolism to enhance

ceramide accumulation; and (iii) delivery of exogenous ceramides to induce

apoptotic signaling (Fig. 1-4).

10

Figure 1-4 Strategies to alter the sphingolipid balance in cancer cells to potentiate

cytotoxicity of chemotherapeutic drugs. 1) Chemotherapy-induced synthesis of pro-apoptotic

sphingolipids or breakdown of pro-survival sphingolipids; 2) Disruption of metabolism of pro-

apoptotic sphingolipids to enhance accumulation; and 3) Nanoscale-based delivery of exogenous

pro-apoptotic sphingolipids in combination with standard chemotherapeutic drugs to induce

apoptotic signaling.

11

Ceramide and induction of cell death

Ample evidence in the literature exists that indicate that ceramide is a

tumor-suppressor lipid and halts tumor progression by inducing cell death and by

cell cycle arrest. Both stimulus-induced intracellular ceramide generation and

exogenous cell-permeable short-chain ceramides induce death in cancer cells by

apoptosis, necrosis or autophagy.

Ceramide and apoptosis

Ceramide generation has been linked to both, the extrinsic and the intrinsic

pathway of apoptosis. Ceramide is an important mediator of initiating cell death

by activation of the pro-apoptotic tumor necrosis factor (TNF) receptor

superfamily, including CD95, TNFR1 and the TNF-related apoptosis-inducing

ligand (TRAIL) receptors TRAILR1 and TRAILR2 [18]. Receptor activation leads

to ceramide synthesis at ceramide-enriched membrane platforms by the de novo

pathway or activation of SMases. The ceramide-enriched membrane platforms

act as scaffolds for localization of pro-apoptotic proteins that initiate intracellular

signaling for cell death, some of which possess ceramide-binding domains.

Ceramide-enriched membrane platforms assemble TRAILR2 and CD95, and act

as a scaffold for the formation of the death-inducing signaling complexes [19-21].

Activation of these cell death receptors induce intracellular ceramide generation

by activating specific CerS and SMases [18]. The importance of ceramide

generation in these cell death pathways has been demonstrated by reports that

inhibition of ceramide synthesis diminishes apoptosis. For instance, pretreatment

with dihydroceramide synthase inhibitor fumonisin B1 (FB1) has been shown to

12

diminish CD95-induced apoptosis in leukemia [22]. Similarly, low ceramide levels

have been correlated to resistance to CD-95 induced apoptosis, TNF-induced

and TRAIL-induced cell death in several cancer models [23-25].

Ceramide also mediates intrinsic pathway-driven apoptosis. Ceramide

accumulation affects mitochondrial bioenergetics and induces conformation of

mitochondrial pro-apoptotic proteins to initiate apoptosis. Intensive research on

the effects of ceramide generation in the mitochondria has convincingly

demonstrated that accumulation of ceramide macrodomains in mitochondria

cause formation of ceramide channels that induce mitochondrial outer membrane

permeabilization (MOMP) [26, 27]. Ceramide accumulation either through

exogenously delivered short chain ceramides or endogenous ceramide

generation synergizes with pro-apoptotic proteins like BAX and BID to

permeabilize mitochondrial membrane and the subsequent release of pro-

apoptotic proteins like cytochrome c and activation of the caspase cascade [28-

31]. Ceramide also initiates intrinsic pathway-driven apoptosis by activating

kinases like p38 MAPK, glycogen synthase kinase (GSK) 3β, JUN N-terminal

kinase (JNK), protein kinase C (PKC) δ or inactivating AKT, which eventually

perturb mitochondrial integrity and cause release of pro-apoptotic proteins [18].

Ceramide is linked to several signaling pathways in apoptosis. The role of

ceramide in inactivating the very crucial AKT pathway in cancer cells has been

widely studied. The downstream targets of ceramide to bring about this

inactivation include PP2A, PKCζ and p38 MAPK [18]. Furthermore, ceramide

activates apoptosis signal-regulating kinase 1 (ASK1) which eventually increases

13

p38 and JNK activation [32]. Additionally, ceramide promotes p53 activation in

certain cancer types which causes a reduction in BAX/BCL-2 ratio, eventually

leading to apoptotic cell death [33, 34]. Finally, ceramide has also been shown to

downregulate anti-apoptotic proteins like survivin to induce apoptotic cell death

[35].

Ceramide and necrosis

The role of ceramide in necrotic cell death is not as well characterized as

apoptosis. Treatment with short chain ceramides like C2 and C6-ceramide

induces necrotic cell death in A20 B-, Raji B- and Jurkat T cells. Cell death was

without caspase-3 activation, DNA fragmentation, cell shrinkage, or chromatin

condensation. FasL-dependent delayed elevation of ceramide promoted caspase

8-driven necrotic morphology after treatment. Inhibition of ceramide production

shifted the mechanism of cell death from necrosis to apoptosis [36]. Further

investigation of the necrotic mechanism has revealed that exogenous C6-

ceramide causes necrosis in lymphoid cells by rapid production of reactive

oxygen species (ROS), loss of mitochondrial membrane potential and ATP

depletion [37]. This is supported by data demonstrating that C2-ceramide-

induced oncotic necrosis in mouse epidermal tumor cells is modulated by a

decline in cellular glutathione and an elevation of ROS [38]. Our lab has also

shown that C6-ceramide delivered as a nanoliposomal formulation induces

necrotic cell death in chronic lymphocytic leukemia (CLL) cells by targeting the

Warburg effect through downregulation of glyceraldehyde 3-phosphate

dehydrogenase (GAPDH) enzyme of the glycolysis pathway and ATP depletion,

14

supporting the report that synthetic ceramides induce non-apoptotic and necrotic

cell death in malignant B-lymphocytes [39, 40]. C2-ceramide has also been show

to predominantly induce necrotic cell death in NB16 neuroblastoma cells.

Although combined treatment with TNFα and cycloheximide is mediated by

intracellular ceramide generation, this stimuli induces apoptosis instead of

necrosis. Thus, C2-ceramide does not faithfully mimic the effects of apoptotic

ligands such as TNFα, which are thought to be mediated by an accumulation of

endogenous ceramide. C2-ceramide targets phosphatidylcholine in these cells

and elicits a mixture of cell death mechanisms, including necrosis and apoptosis,

the former being more predominant [41]. Ceramide also induces non-apoptotic,

caspase-independent cell death by inducing ROS generation in A172 human

glioma cells. NF-kappaB is involved in the regulation of ceramide-induced cell

death in human glioma cells [40]. Lastly, it has also been observed that certain

cellular parameters play an important role in determining the mechanism of cell

death after ceramide treatment. For instance, in Hep-G2 cells the mitochondrial

respiratory function determines the mechanism of cell death after treatment with

exogenous short chain ceramides. Herein, C2-ceramide induced necrosis which

was a result of 80% inhibition of the mitochondrial respiratory function leading to

ATP depletion and ROS generation. In contrast, C6-ceramide induced apoptotic

cell death in the same cells since mitochondrial function was not inhibited and

ATP production not diminished [42].

15

Ceramide and autophagy

Despite ceramide being a promoter of autophagy, its role in mediating

autophagic cell death is confounded by the role of autophagy in cancer cells, i.e.

lethal versus survival autophagy. Increased levels of long-chain ceramides

(C14:0 – C20:0 ceramides) and especially dihydroceramides have been

associated with both lethal and survival autophagy in different cancer cell types

[43, 44]. It is believed that the fate of the autolysosome dictates the function of

autophagy as lethal versus survival. Intracellular generation of sphingosine and

S1P in the autolysosomes by the hydrolysis of dihydroceramides and ceramides

promotes pro-survival autophagy. In contrast, accumulation of dihydroceramides

in the autolysosomes can enhance autolysosomal membrane permeability and

cause the release of cathepsins, thereby causing apoptotic cell death [45]. In this

case, ceramides promote lethal autophagy in cancer cells.

Thus, as discussed earlier sphingolipid-based therapeutics aim to alter the

sphingolipid balance to induce cell death by either elevating pro-apoptotic

sphingolipids, especially ceramide, or down-regulating pro-survival sphingolipids

such as S1P. This can be achieved by: (i) chemotherapy-induced synthesis of

pro-apoptotic ceramides or breakdown of pro-survival sphingolipids; (ii) disruption

of ceramide metabolism to enhance ceramide accumulation; and (iii) delivery of

exogenous ceramides to induce apoptotic signaling. The next section discusses

the first strategy as a proof-of-concept to demonstrate that the effectiveness of

current chemotherapies and investigational drugs is partially or completely

16

mediated by endogenous ceramide generation. The discussion will then focus on

the development of exogenous ceramide-based therapeutics for cancer.

Chemotherapy-induced ceramide generation by the de novo pathway

Many chemotherapeutics increase ceramide levels by upregulating the de

novo pathway of ceramide synthesis. This is accomplished by increasing the

activity of serine palmitoyl transferase (SPT), ceramide synthase (CerS), or both.

There are six CerS isoforms (CerS1-6), which are also known as longevity

assurance (LASS) genes, with each isoform corresponding to specific resulting

carbon chain lengths of ceramide [46]. CerS1 produces C18 ceramide, CerS2

C20-C26, CerS3 C22-C26, CerS4 C18 and C20, CerS5 C16, and CerS6 C14

and C16 ceramide [46]. Several inhibitors are routinely used to study the de

novo pathway, including fumonisin B1 (FB1), which inhibits CerS; and L-

cycloserine and myriocin, which inhibit SPT.

The taxanes docetaxel and paclitaxel are two chemotherapeutic agents

that increase ceramide levels by the de novo pathway. Both of these drugs act

by preventing microtubule disassembly, leading to cell cycle arrest [47].

Docetaxel, which is used to treat metastatic ovarian, breast, lung, head and neck,

and prostate cancer, increases CerS1 and CerS2 levels, while decreasing

sphingosine kinase 1 (SK-1) levels in prostate cancer cells [48]. Ceramide is also

a critical regulator of taxane-induced cell death since overexpression of the

ceramide transfer protein, CERT, reduces the sensitivity of cells to taxanes [49].

Similarly, paclitaxel-induced apoptosis in breast cancer cells is dependent on

ceramide produced by the de novo pathway [50].

17

In addition to the taxanes, the anthracyclines doxorubicin and

daunorubicin also affect ceramide metabolism in tumors. In addition to its effects

on DNA, doxorubicin also causes a CerS-dependent increase in ceramide levels

in neuroepithelioma and neuroblastoma cells [51]. However, inhibition of de

novo ceramide generation using FB1 surprisingly has no effect on doxorubicin-

induced apoptosis in these cells [51]. On the other hand, in head and neck

cancer cells, doxorubicin alone or in combination with gemcitabine causes

apoptosis that is dependent on CerS1 and the generation of C18:0 ceramide

[52]. This was confirmed in an animal model of head and neck cancer in which

doxorubicin in combination with gemcitabine decreased C16:0 ceramide in the

tumor, but increased intratumoral CerS1 and C18:0 ceramide [52]. Similar to

doxorubicin, camptothecin also exerts antitumor properties in follicular thyroid

carcinoma through ceramide accumulation via de novo synthesis. Both of these

drugs cause activation of ceramide synthesis without any effects on SMases.

Apoptosis is mediated by ceramide elevation and can be enhanced by the use of

inhibitors of ceramide clearance [53]. Similar to doxorubicin, daunorubicin also

acts by intercalating into DNA [47] and induces apoptosis in leukemia cells by de

novo ceramide generation and ceramide generation by the action of SMase [17].

It is used to treat acute myeloid leukemia (AML) and AIDS-related Kaposi’s

sarcoma (liposomal formulation) [47].

Besides the taxanes and the anthracyclines, vorinostat, a histone

deacetylase inhibitor in combination with sorafenib, a kinase inhibitor also

modulates ceramide metabolism. Vorinostat in combination with sorafenib

18

causes de novo pathway-dependent ROS production and cell death in HCC cells

[54]. This is accompanied by an increase in C16 ceramide, as well as C16, C18,

C22, C24:0, and C24:1 dihydroceramide [54].

Another chemotherapeutic agent that increases ceramide levels is

fludarabine, an inhibitor of DNA synthesis that is used to treat chronic

lymphocytic leukemia (CLL) [47]. Fludarabine causes a de novo pathway-

dependent 2.5- to 3 fold elevation in ceramide levels in CLL cells 6 hours after

treatment. Pretreatment with fumonisin B1 significantly rescues fludarabine-

induced ceramide generation and apoptosis [55]. In addition, CLL cells treated

with non-physiological, short chain C6 ceramide undergo apoptosis and necrosis

[39, 55], suggesting that fludarabine may kill CLL cells via upregulation of

ceramide.

Certain chemotherapeutic drugs generate ceramide by acting on SPT.

Etoposide-induced apoptosis in human leukemia cells is mediated by ceramide

synthesized by the activation of SPT. Ceramide formed by this pathway has

distinct functions in this model system as compared to that formed by the SMase

pathway. Ceramide synthesized by the de novo pathway in this model is not

involved in caspase-induced poly (ADP-ribose) polymerase (PARP) cleavage but

instead perturbs membrane integrity [56].

In addition to current chemotherapies, there are drugs approved by the

United States Food and Drug Administration (FDA) for other uses that are

currently being investigated as potential anti-cancer agents. One of these is the

19

COX-2 inhibitor celecoxib, which is currently used to treat pain, inflammation, and

arthritis, and to prevent polyposis coli [47]. Recent work indicates that celecoxib

decreases cell viability in colorectal carcinoma cells in a CerS6-dependent

manner [57]. This is accompanied by an increase in sphingosine,

dihydrosphingosine, C14:0 ceramide, C16:0 ceramide and C18:0 ceramide, and

a decrease in C24:0 ceramide [57]. Furthermore, C16:0 ceramide is found at

elevated levels in tumors treated with celecoxib in an in vivo model of colorectal

carcinoma [57]. A number of investigational chemotherapeutics also upregulate

ceramide via the de novo pathway. Investigational drugs such as Valspodar,

inostamycin, and spisulosine induce apoptosis in a variety of cancer cell types

via de novo ceramide generation [58-62]. More recently, the endocannabinoid

analog R (+)-methanandamide (RMA) has been shown to increase C16 ceramide

levels in neuroglioma cells via the de novo pathway [63]. RMA, which increases

CerS3 and CerS6 in mantle cell lymphoma (MCL) cells, also increases C16, C24,

and C24:1 ceramide levels via the de novo pathway in MCL cells [64]. This

increase is functionally significant, as RMA-induced cell death in these cells is

also dependent on the de novo pathway [64].

In addition to RMA, another investigational drug that increases ceramide

levels is fenretinide (4-HPR). It is a synthetic retinoid N-(4-hydroxyphenyl) that

induces cell death in various cancers like neuroblastoma [65] and in cell lines

from cervical carcinoma [66] and acute myeloid leukemia (AML) [67]. Studies

have shown that the drug elevates intracellular ceramide levels via de novo

synthesis by increasing the activity of both SPT and CerS [68], and induces p53-

20

and caspase-independent apoptosis in neoplastic cells [69]. 4-HPR also

functions as a dose-dependent inhibitor of ceramide destaurase [70], indicating

that the cytotoxic sphingolipid species may be dihydroceramide or

dihydrosphingosine rather than ceramide [71]. Moreover, 4-HPR-induced

cytotoxicity is synergistic with inhibitors of ceramide metabolism like sphingosine

kinase inhibitors and glucosylceramide synthase inhibitors [72]. Other effects

induced by 4-HPR include generation of ROS and enhanced expression of LC3B

(form II). 4-HPR is currently in clinical trials for ovarian cancer, neuroblastoma,

lymphoma and leukemia [73].

Finally, the investigational AMPK inhibitor Compound C increases CerS5,

C16:0 ceramide, and C18:0 ceramide, and decreases in sphingosine in breast

cancer cells [74]. Clearly, many current and investigational chemotherapeutics

increase ceramide levels by targeting the de novo pathway. This increase in

ceramide, in turn, induces apoptosis in cancer cells.

Chemotherapy-induced ceramide generation by sphingomyelinase pathway

Independent of the de novo pathway, ceramide can also be generated

when SMase hydrolyzes sphingomyelin to produce ceramide. A number of

chemotherapeutics increase ceramide levels in cancer cells by increasing the

activity of SMase. Some drugs, such as daunorubicin, upregulate ceramide by

modulating both the de novo pathway and SMase activity. Besides upregulating

the de novo pathway, it also increases SMase activity in leukemia cells [75]. In

addition, in breast cancer cells, daunorubicin enhances binding of the

transcription factor Sp1 to the nSMase2 gene, leading to increased nSMase2

21

levels and a nSMase-dependent decrease in cell viability [76]. It should be

noted that daunorubicin is currently not approved for the treatment of breast

cancer.

Gemcitabine is another drug that targets both the de novo pathway and

SMase. It increases aSMase activity in pancreatic cancer cells [77]. In addition,

gemcitabine increases aSMase activity in glioma cells, increasing the levels of

C16 and C24 ceramide and an aSMase-dependent decrease in cell survival [78].

One chemotherapeutic that elevates ceramide levels via SMase

independent of the de novo pathway is cytarabine, which increases nSMase

activity and ceramide levels in acute myelogenous leukemia (AML) cells [79].

Cisplatin is another drug that affects SMase activity. Cisplatin, which is a

platinum coordination complex that causes DNA cross-links leading to cell death,

is used to treat testicular, ovarian, bladder, head and neck, cervical, endometrial,

lung, anal, rectal, esophageal, and central nervous system (CNS) cancer, as well

as neoplasms of childhood [47]. It increases ceramide levels and causes

apoptosis in glioma cells in a nSMase-dependent manner, wherein it transiently

increases aSMase activity and causes it to be redistributed to the plasma

membrane [80]. Additionally, a cisplastin-induced increase in SMase activity and

the subsequent accumulation of ceramide levels are essential for cytoskeletal

remodeling following treatment with cisplastin, such as loss of

lamellipodia/filopodia and dephosphorylation and redistribution of the actin-

binding protein ezrin [81].

22

Another chemotherapeutic that is currently used in the clinic and that

affects SMase is rituximab, which is a monoclonal antibody to CD20 that is used

to treat lymphoma and chronic lymphocytic leukemia [47]. This drug increases

aSMase activity in lipid rafts, increasing ceramide levels in B-lymphoma cells,

thereby inhibiting cell growth in these cells [82]. Furthermore, exogenous

treatment with C16 ceramide decreases cell viability in this system [82]. Taken

together, these findings indicate that rituximab may inhibit B-lymphoma cell

proliferation by activating aSMase, thus increasing ceramide levels.

Recently, the investigational drug stichoposide C, a marine triterpine

glycoside, has been shown to cause apoptosis in leukemia and colorectal cancer

cells in an aSMase- and nSMase-dependent manner [83]. It also inhibits tumor

growth in mouse models of leukemia and colorectal cancer [83]. In addition,

stichoposide C-treated tumors contained elevated levels of ceramide [83].

Betuletol 3-methyl ether, a natural phenylbenzo-γ-pyrone, is another

investigational drug that increases ceramide levels in cancer cells [84]. It causes

apoptosis in leukemia cells and increases aSMase activity and ceramide levels

[84]. Finally, the investigational drug withanolide D acts by increasing nSMase

activity, leading to increased ceramide levels and apoptosis in leukemia cells

[85]. It decreases cell viability in leukemia cells but not normal lymphocytes [85].

Furthermore, it induces apoptosis in primary cells from both myeloid and

lymphoid leukemia patients [85]. Withanolide D is a good example of an

investigational therapy that modulates SMase. While several current

23

chemotherapies target SMase, even more such drugs could be added to the

oncology arsenal in the future.

Nanotechnology-based drug delivery of ceramide

Delivering exogenous ceramides to cancer cells is another strategy to

perturb sphingolipid levels for inducing apoptosis. However, this strategy is

greatly limited by the hydrophobicity and insolubility of ceramide molecules,

which in turn restricts delivery in cell culture systems and in vivo. Soluble

ceramide analogs have been developed to circumvent this problem. Short chain

ceramides like C2-, C6- and C8-ceramides are cytotoxic in multiple cancer

models [86]. Chemical modifications of short-chain ceramides improve their

solubility, permeability and pharmacokinetics. Certain chemical modifications that

have been tested in cancer models include uracil-linked ceramides [87],

serinamides [88], serinols [89, 90] and 4,6-diene-ceramides [91]. In addition,

cationic water-soluble pyridinium-ceramides have been developed which

preferentially accumulate in mitochondria and induce cell death by mitochondrial

permeabilization and Bax translocation [92]. These are effective in inducing cell

death in head and neck squamous cell cancer [93] and breast cancer cell lines

[94]. These analogs also synergize with gemcitabine to cause cell cycle arrest in

G0/G1 phase, retard growth and inhibit telomerase activity in human head and

neck squamous cell carcinomas in vitro and in vivo [93, 95]. Other structural

analogs of ceramide that have shown selective cytotoxicity in drug-resistant

human breast cancer cells compared to normal breast epithelial cells include 5R-

OH-3E-C8 ceramide, benzene-C4-ceramide and adamantyl-ceramide [96].

24

Additional novel ceramide analogs including AD2646, AD2672, AD2665, AD2646

and AD2687 have cytotoxic effects on leukemic cells [97].

Ceramide-based therapies face challenges like high insolubility and

difficulties to design formulations. Nanoscale-based formulations have thus been

developed and investigated to deliver these therapies. Nanoemulsions,

nanoliposomes, calcium phosphosilicate nanoparticles and biodegradable linear-

dendritic nanoparticles are used for delivering ceramide-based therapeutics [98].

Ceramide delivered via nanoscale formulations has been shown to induce cell

death selectively in cancer cells while sparing normal cells [98].

Novel oil-in-water nanoemulsions have been evaluated as delivery

vehicles for potential combination therapies in vitro. EGFR-targeted

nanoemulsions containing myrisplatin and C6-ceramide show synergistic in vitro

cytotoxicity in ovarian cancer cells and also possess potential diagnostic

capabilities [99]. Similarly, coadministration of paclitaxel and ceramide in

nanoemulsion formulations induces enhanced cytotoxicity and apoptotic activity

in human glioblastoma cells in vitro [100]. Sustained release of C6-ceramide from

thermoresponsive and biodegradable linear-dendritic nanoparticles induces

apoptosis in breast adenocarcinoma cells in vitro with hyperthermia, thus

presenting a promising formulation to deliver bioactive sphingolipids for treatment

of solid tumors in conjunction with hyperthermia [101]. Biodegradable polymeric

nanoparticles have also been evaluated for modulation of drug resistance in

cancer cells in vitro and in vivo. Paclitaxel and tamoxifen administered in

biodegradable poly (ethylene oxide) - modified poly (epsilon-caprolactone) (PEO-

25

PCL) nanoparticles possess significant antitumor efficacy in ovarian carcinoma in

vitro and in vivo. Tamoxifen in the formulation reverses drug resistance by

inhibiting P-gp and GCS, thus elevating intracellular ceramide levels in cancer

cells [102]. Paclitaxel and C6-ceramide PEO-PCL nanoparticles have also been

used to chemosensitize resistant human ovarian cancer cell lines to induce

apoptotic cell death [103] and suppress growth in xenograft tumor models [104].

Additionally, we have demonstrated the cytotoxic effects of C10-ceramide-loaded

calcium phosphate nanocomposite particles in drug-sensitive and drug-resistant

breast cancer and melanoma cells in vitro [105].

Encapsulation of chemotherapeutic drugs into nanoliposomes has been a

largely successful delivery formulation in cancer models in vitro and in vivo. Our

lab has extensively studied nanoliposomes as a suitable delivery formulation for

ceramides in several cancer models. We have shown that nanoliposomes loaded

with short-chain ceramides suppress tumor growth in models of breast cancer

[106, 107], J774 sarcoma [108], melanoma [109], hepatocellular carcinoma [110],

large granular lymphocytic (LGL) leukemia [35], chronic lymphocytic leukemia

(CLL) [39], natural killer cell leukemia [111] and pancreatic cancer [112].

Mechanistically, the targets of nanoliposomal C6-ceramide include survivin in

LGL leukemia [35], GAPDH in CLL [39] and AKT/PKB and Erk in breast cancer

[107], hepatocellular carcinoma [110], pancreatic cancer [112] and melanoma

models [109]. Our studies and extensive in vivo toxicology studies by National

Cancer Institute sponsored agency, Nanotechnology Characterization Laboratory

(http://ncl.cancer.gov/working_technical_reports.asp) have confirmed that C6-

26

ceramide nanoliposomes have minimum adverse toxicity and elicit apoptosis

selectively in cancer cells [113]. We have established that selectivity of C6-

ceramide nanoliposomes for cancer cells can be attributed to the inhibitory action

of ceramide on the Warburg effect prevalent in cancer cells [39].

Our lab has also evaluated C6-ceramide nanoliposomes as a platform for

combinatorial therapy with other neoplastic agents. We have shown that C6-

ceramide nanoliposomes synergize with encapsulated sorafenib to reduce tumor

development in melanoma and breast cancer cells [109], synergize with

gemcitabine or liposomal PDMP to exhibit antitumor effects on pancreatic tumor

xenografts [112] and synergize with PPMP to induce apoptosis in natural killer

cell leukemia [111]. Recently, in collaboration with the Cabot group we showed

that nanoliposomes loaded with C6-ceramide and tamoxifen served as a

promising regimen for refractory breast cancers like the triple-negative breast

cancer [114]. Tamoxifen amplifies C6-ceramide-induced cytotoxicity in the triple-

negative breast cancer cells by multiple effects like cell cycle arrest, lysosomal

membrane permeability and inhibition of acid ceramidase [114]. C6-ceramide

nanoliposomes also exhibit synergy with the autophagy inhibitor vinblastine to

induce apoptotic cell death in vitro and in vivo in hepatocarcinoma and colorectal

cancer models, potentially mediated by an autophagy mechanism [115]. In

addition to chemotherapeutic drugs, multidrug resistance modulators are also

favorable adjuvants for C6-ceramide nanoliposomes. This strategy is justified by

studies reporting that resistance to C6-ceramide cytotoxicity is a result of

expression of P-gp in some cancer cells [116]. P-gp inhibitors also alter

27

sphingolipid levels in cancer cells. For instance, the multidrug resistance

modulator SDZ PSC 833 elevates intracellular ceramide levels by inducing the

de novo pathway [117]. P-gp antagonists like tamoxifen, verapamil, and

cyclosporine A can also be used in conjunction with cytotoxic drugs like

doxorubicin to decrease GlcCer, subsequently increasing ceramide levels and

sensitizing cells to cytotoxic drugs [59, 118]. In collaboration with the Cabot

group, we have shown that C6-ceramide nanoliposomes-mediated cytotoxicity in

cancer cell lines can be augmented by P-gp antagonists like tamoxifen,

verapamil and VX-710 [119, 120]. C6-ceramide and tamoxifen induced apoptotic

cell death in colorectal cancer cells characterized by PARP cleavage,

mitochondrial membrane permeabilization, caspase-dependent apoptosis and

G1/G2 cell cycle arrest. Moreover, the combinatorial treatment exhibited synergy

and induced upregulation of tumor suppressor p53 [119].

Co-administration of paclitaxel and C6-ceramide exhibit synergy to induce

cytotoxicity in pancreatic cancer cells via transient activation of EGFR and ERK

pathway [121] and ovarian cancer cells [122]. An interesting study revealed that

in the absence of paclitaxel, exogenous C6-ceramide enters the cell through a

predetermined initiation site of mitosis, or diffuses into cells through water

channels and caveolae-mediated endocytosis [122]. However, the combination

induces synergistic cytotoxicity in cancer cells as paclitaxel disrupts cytoskeletal

proteins, thus enabling an even distribution of C6-ceramide in the cytoplasm of

the cells [122]. Other reports have demonstrated that C6-ceramide also

synergizes with histone deacetylase inhibitors like trichostatin A to display

28

anticancer effects in in vivo mice xenograft pancreatic and ovarian cancer

models [123]. The authors have delineated the mechanism of this synergy and

have demonstrated PP1-mediated inactivation of Akt/mTOR and increased α-

tubulin acetylation as events causing cancer cell death. Furthermore, the

combination resulted in a very pronounced elevation in intracellular ceramide

levels and induction of cell death in cancer cells [123]. C6-ceramide also

synergizes in inducing apoptotic cell death in leukemia cells with other neoplastic

agents like the cationic peptide, bovine lactoferricin [124].

In conclusion, Fig. 1-5 summarizes the points of intervention in the

sphingolipid pathway and the drugs being studied for developing sphingolipid-

based cancer therapeutics.

29

Figure 1-5 Points of intervention in the sphingolipid pathway. Enzymes catalyzing ceramide

synthesis or ceramide metabolism can be activated or inhibited respectively to cause ceramide

accumulation and induce death in cancer cells. SPT, serine palmitoyl transferase; CerS,

ceramide synthases; DDase, dihydroceramide desaturase; SMase, sphingomyelinase; SMS,

sphingomyelin synthase; SphK, sphigosine kinase; GCS, glucosyl ceramide synthase; PPMP, 1-

phenyl-2-palmitoylamino-3-morpholino-1-propanol; PDMP, 1-phenyl-2-decanoylamino-3-

morpholino-propanol; 4‑HPR, N‑(4‑hydroxyphenyl) retinamide; DHS, L-threo-dihydrosphingosine;

SKI II, 2-(p-hydroxyanilino)-4-(p-chlorophenyl) thiazole.

30

Chronic Lymphocytic Leukemia (CLL)

CLL is the most prevalent form of adult leukemia in Western countries. As

per 2015 statistics, CLL accounts for approximately 30% of total leukemia cases

diagnosed and it accounts for 20% of deaths from all kinds of leukemia [125].

CLL mainly affects older adults. The average age at the time of diagnosis is

around 71 years. It is rarely seen in people under age 40, and is extremely rare

in children.

CLL is a malignant lymphoproliferative disorder of mature B lymphocytes.

The disease is characterized by an accumulation of mature looking B

lymphocytes in the blood, bone marrow, lymph nodes or other lymphoid tissues.

Leukemic B cells express characteristic surface markers consisting of CD19,

CD20 (weak) and CD23, with co-expression of CD5. Most CLL cells are arrested

in the G0/G1 phase and are highly resistant to apoptosis, eventually leading to

an accumulation of malignant cells [126]. A large body of work has demonstrated

that several factors like microenvironmental stimuli, antigenic drive and

epigenetic and genetic deregulation dictate the pathogenesis of CLL.

CLL is classified into different sub-types that determine the prognosis of

the disease and the treatment strategy. One classification is based on the degree

of somatic hypermutation, i.e. whether the cells express mutated or unmutated

immunoglobulin heavy chain variable region (IGHV) genes. The two groups

follow a different clinical course, with poorer survival in patients exhibiting

unmutated IGHVs. In addition to this determinant, approximately 80% of CLL

cases also show chromosomal aberrations. These genetic aberrations are

31

observed in both IGHV mutated and unmutated CLL, the latter being associated

with higher incidence of high-risk aberrations [125]. Deletion in band 13q14 is the

most common genetic aberration in CLL and has a favorable prognosis. This part

of the chromosome contains mir-15a and mir-16-1 which have been implicated in

CLL pathogenesis [127]. Trisomy 12 is another frequent chromosomal

abnormality in CLL, however, the corresponding molecular implications on

pathogenesis remains unknown. Deletion in band 11q23 is not a very frequent

aberration in early stage disease, but is associated with rapid disease

progression and excessive lymphadenopathy [128, 129]. This deletion usually

corresponds with deletion of the ataxiatelangiectasia-mutated (ATM) gene, which

is an essential component of the cell cycle checkpoint system. Deletion in the

ATM gene is characterized by extreme sensitivity to irradiation, genomic

instability and a predisposition to lymphoid malignancies [125]. Lastly, deletion in

band 17p13, corresponding to p53 deletion and TP53 mutation is associated with

poor prognosis [125].

The microenvironment in the lymphoid organs plays a crucial role in the

pathogenesis of CLL. The microenvironment consists of T-cells, stromal cells and

soluble factors. Soluble factors from the microenvironment provide CLL cells with

a protective environment and provide anti-apoptotic and pro-proliferative stimuli

that are necessary for the progression of the disease. For instance, CLL cells

recruit CD3+ T cells which also express CD40L and CD4. CD40L from these

accessory cells initiate the indispensable B-cell receptor (BCR) signaling in CLL

cells after interaction with CD40 receptor. Such stimuli induce production of anti-

32

apoptotic proteins like survivin, Mcl-1 and Bcl-2 in CLL cells which are crucial for

cell viability [125, 130]. Furthermore, the stromal microenvironment promotes a

metabolic switch in CLL cells from mitochondrial respiration to aerobic glycolysis,

thus conferring growth advantage and conferring chemoresistance to CLL cells

[131]. In conclusion, ample evidence in the literature exists emphasizing the

crucial role of stromal microenvironment in the pathogenesis and

chemoresistance of CLL. A multitude of molecules including integrins, spleen

tyrosine kinase, stromal derived factor-1, Notch, CD44, and thioredoxin have

been identified to be part of the stromal cross talk [125].

The standard treatment regimen for physically fit CLL patients includes a

combination chemo-immunotherapy with fludarabine, cyclophosphamide and

rituximab with an overall response rate of approximately 90% and complete

remission of 72% [132, 133]. A combination of purine analogs, alkylating agents,

monoclonal antibodies (immunotherapy) and BCR pathway and tyrosine kinase

inhibitors is the standard treatment. Novel drugs recently incorporated in the

treatment regimen include ibrutinib, which targets an important component of

BCR signaling, Bruton’s tyrosine kinase (BTK). Another novel orally available

agent is idelalisib, a phosphatidylinositol-3-kinase (PI3K) δ inhibitor. Both of

these drugs have been for relapsed/refractory disease and first-line treatment of

patients with TP53 mutation/deletion [134]. Unfortunately these advances do not

benefit older CLL patients due to their frail health [135]. Overall, CLL is incurable

with the current therapies, with allogeneic stem cell transplantation as the only

potentially curative treatment option in CLL. However, this option is also limited to

33

young and relatively healthy patients. Despite these advances in therapeutics,

eventual drug resistance and relapse ultimately cause CLL to be an incurable

and chronic disease [136]. Further research is needed to develop therapeutic

strategies.

The role of sphingolipids in CLL pathogenesis or treatment has not been

explored yet. Synthetic ceramides have been shown to induce non-apoptotic and

necrotic cell death in malignant B-lymphocytes [137]. Additionally, a few reports

have also demonstrated that sphingolipids mediate cell death in CLL cells. It has

been reported that membrane microdomain sphingolipids are required for anti-

CD20-induced cell death in CLL cells [138]. The authors showed that resistance

to anti-CD20-induced cell death is associated with a defective recruitment of Csk-

binding protein, resulting in a lack of sphingomyelin and ganglioside M1 at the

outer leaflet of the plasma membrane of malignant B cells. Inducing P-

glycoprotein in resistant cells restored sensitivity to anti-CD20 antibody as the

inducer normalized the quantity of sphingomyelin within the membrane [138].

Another recent report has uncovered a novel link between BCR signaling and

sphingolipid metabolism. It has been reported that BCR controls

chemoresistance of primary CLL cells by controlling glucosylation of ceramides

[139]. Specifically, BCR engagement increases levels of anti-apoptotic

glucosylated ceramides via upregulation of UDP-glucose ceramide

glucosyltransferase (UGCG), an enzyme which converts pro-apoptotic ceramide

to anti-apoptotic glucosylceramide. The authors have shown that inhibitors of

BCR signaling sensitize resistant CLL cells towards ABT-737 drug via UGCG

34

inhibition [139]. Another report has demonstrated that sphingolipid metabolism is

a potential novel mechanism of CLL [140]. Taken together, these reports provide

compelling evidence in support of the use of sphingolipid-modulating strategies,

and more specifically, ceramide-based strategies as novel therapeutics in CLL.

Conclusions

The role of nanoliposomal C6-ceramide in inducing cell death in several

types of solid and non-solid cancers is well understood. However, the effect of

nanoliposomal C6-ceramide treatment in CLL remains unclear. This dissertation

primarily focuses on delineating the molecular mechanisms of C6-ceramide-

induced cell death in CLL. Identifying the key signaling pathways inducing cell

death after ceramide treatment is also valuable to uncover additional targets for

potential combination therapies with ceramide nanoliposomes. Encapsulation of

chemotherapeutic drugs into nanoliposomes has been a largely successful

delivery formulation in cancer models in vitro and in vivo. Moreover, the ongoing

success of using ceramide nanoliposomes as a platform for combinatorial

therapy with other neoplastic agents presents a promising future to this endeavor

of developing more effective therapeutics for CLL.

35

CHAPTER 2: Nanoliposomal C6-ceramide target the Warburg effect in chronic lymphocytic leukemia

I would like to recognize that Chapter 2 of my dissertation titled “C6-ceramide

nanoliposomses target the Warburg effect in chronic lymphocytic leukemia” is

derived from the following published literature:

Ryland LK*, Doshi UA*, Shanmugavelandy SS, Fox TE, Aliaga C, Broeg K, Baab KT, Young M, Khan O, Haakenson JK, Jarbadan NR, Liao J, Wang HG, Feith DJ, Loughran TP Jr, Liu X, Kester M. C6-ceramide nanoliposomes target the Warburg effect in chronic lymphocytic leukemia. PLoS One. 2013 Dec 19;8(12):e84648. I am the co-first author for this published work.

This project was underway when I joined the Kester lab in 2011. Dr. Lindsay

Ryland and I contributed towards the design of the experiments within this

particular chapter. Specifically, I was a major contributor towards the following

figures: Fig 2-1A, Fig. 2-3A, Fig. 2-4A and Fig 2-5. I am in no way attempting to

claim intellectual property over the design of all the experiments within this

particular chapter. Therefore, I would like to acknowledge Dr. Lindsay Ryland

and the other contributing authors for the work presented in this chapter.

36

Abstract

Ceramide is a sphingolipid metabolite that induces cancer cell death.

When C6-ceramide is encapsulated in a nanoliposome bilayer formulation, cell

death is selectively induced in tumor models. However, the mechanism

underlying this selectivity is unknown. As most tumors exhibit a preferential

switch to glycolysis, as described in the “Warburg effect”, we hypothesize that

ceramide nanoliposomes selectively target this glycolytic pathway in cancer. We

utilize chronic lymphocytic leukemia (CLL) as a cancer model, which has an

increased dependency on glycolysis. In CLL cells, we demonstrate that C6-

ceramide nanoliposomes, but not control nanoliposomes, induce caspase 3/7-

independent necrotic cell death. Nanoliposomal ceramide inhibits both the RNA

and protein expression of GAPDH, an enzyme in the glycolytic pathway, which is

overexpressed in CLL. To confirm that ceramide targets GAPDH, we

demonstrate that downregulation of GAPDH potentiates the decrease in ATP

after ceramide treatment and exogenous pyruvate treatment as well as GAPDH

overexpression partially rescues ceramide-induced necrosis. Finally, an in vivo

murine model of CLL shows that nanoliposomal C6-ceramide treatment elicits

tumor regression, concomitant with GAPDH downregulation. We conclude that

selective inhibition of the glycolytic pathway in CLL cells with nanoliposomal C6-

ceramide could potentially be an effective therapy for leukemia by targeting the

Warburg effect.

37

Introduction

Sphingolipids are a class of complex cellular lipids that serve both a

structural role in the cellular membrane as well as an intracellular signaling role

within the cell. Several types of sphingolipid metabolites have been shown to

influence the balance between mitogenesis and apoptosis. Of particular interest

is the sphingolipid metabolite, ceramide, which regulates differentiation,

senescence and cell cycle arrest. Induction of cell death by this endogenous

lipid-derived second messenger occurs either via apoptotic, autophagic, or

necrotic cell death pathways [41, 141, 142]. Ceramide inhibits cell proliferation

and induces apoptosis via mechanisms such as dephosphorylation and/or

inactivation of molecules including Akt, phospholipase D, ERK, Bcl-2, survivin,

PKC-α, and pRB [4, 143, 144], as well as activation of JNK kinases [4, 145], or

PKC zeta which, results in suppression of Akt-dependent mitogenesis [146].

Therefore, it is not surprising that dysregulated ceramide metabolism and

signaling has been linked to a variety of human diseases, including cancer.

Based on its ability to selectively block tumor initiation and metastasis, ceramide

has been termed the ‘tumor-suppressor lipid’ [4]. Many cancer chemotherapies

have been shown to generate endogenous ceramide, and when de novo

generation of ceramide is inhibited, the cellular response to cytotoxic

chemotherapeutic agents decreases [4]. In addition the accumulation of

endogenous ceramides or exogenous ceramide treatment is more toxic to tumor

cells than to normal cells [144, 147]. However, the exact mechanism of selectivity

is unknown.

38

One proposed mechanism for how ceramide mediates cell death induction

is through downregulation of nutrient transporter proteins possibly via nutrient

deprivation. [148]. As cancer cells have an increased dependence on glucose,

these nutrient transporters and/or the glycolytic pathway are typically

upregulated. One hallmark of cancer cells is their ability to avidly take up

glucose and convert it to lactate, even in the presence of sufficient oxygen.

Deemed the “Warburg effect,” this altered glycolytic dependency favors less

efficient generation of ATP compared to the oxidative phosphorylation process

which occurs in normal cells [149, 150]. Many human cancers display increased

levels of glycolytic enzymes compared to normal tissue [151]. Consequently, a

variety of chemotherapeutic glycolytic inhibitors or PET modalities are currently

under investigation as potential “Warburg-targeted” therapeutic or diagnostic

imaging tools [152, 153]. Recently, the role of sphingosine kinases in regulating

the Warburg effect in prostate cancer cells has been reported [154]. Treatment of

LNCaP prostate cancer cells with SKi, a non-selective sphingosine kinase

inhibitor, increases intracellular levels of ceramide and sphingosine and indirectly

antagonizes the Warburg effect, resulting in apoptosis of LNCaP cells.

Chronic lymphocytic leukemia (CLL) is the most common B-cell

malignancy in the Western world which presently has no known curative therapy

[155]. Previous studies have demonstrated that treatment with exogenous short-

chain C2-ceramide results in induction of cell death in malignant cells isolated

from CLL patients [137]. Recent advances in nanotechnology have illustrated

the feasibility of generating nanoliposomes that encapsulate hydrophobic

39

compounds, like ceramide, to facilitate treatment of CLL. While it is understood

how nanoliposomal ceramide induces cell death in several types of cancers and

hematological malignancies, the effect of nanoliposomal ceramide treatment in

CLL remains unclear. Currently, several nanoliposomal formulations of anti-

cancer drugs have been approved by the FDA and are the standard of care

[156]. For instance, the efficacy of fludarabine, the cancer chemotherapy

commonly used to treat CLL patients, and which acts via intracellular ceramide

accumulation, is enhanced after being encapsulated in nanoliposomes [157,

158]. Our laboratory has demonstrated that encapsulation of ceramide in a

nanoliposome versus non-liposomal organic formulations increases the cytotoxic

potential with significant less toxicity [159]. Our laboratory has also demonstrated

that the short chain C6-ceramide nanoliposomal formulation displays anti-

proliferative effects in vitro, as well as results in tumor regression in several

animal models of cancer [110, 144, 147, 159].

In this study we sought to elucidate the effect of nanoliposomal C6-

ceramide treatment in CLL. Our data suggest that this treatment is targeting

glucose utilization and results in activation of a caspase-independent, necrotic

cell death mechanism. We have identified glyceraldehyde 3-phosphate

dehydrogenase (GAPDH) as a novel target of ceramide in CLL cells. In the

current study, we conclude that ceramide targets the Warburg effect in cancer

cells and selectively induces necrotic cell death in CLL.

40

Materials and Methods

Reagents

Antibodies specific for caspase 3, poly ADP ribose polymerase (PARP), GAPDH,

-actin and α-tubulin were purchased from Cell Signaling Technology Inc.

(Beverly, MA), glucose transporter 1 (GLUT1) antibody from Abcam (Cambridge,

MA) and lactate degydrogenase (LDH) antibody from Epitomics (Burlingame,

CA). For Western blotting, 12% precasted Nupage electrophoresis gels from

Invitrogen (Carlsbad, CA), and chemiluminescence reagent from Amersham

Biosciences Inc. (Piscataway, NJ) were obtained. Other reagents include zVAD-

fmk and pyruvate from Sigma (St. Louis, MO), dasatinib from Toronto Research

Chemicals Inc. (Ontario, Canada) and 3-bromopyruvate from Enzo Life Sciences

(Farmingdale, NY).

Patient characteristics and preparation of peripheral blood mononuclear

cells

All patients met the clinical criteria of CLL and were not on treatment at the time

of sample acquisition. Peripheral blood specimens from CLL patients were

obtained and informed consents signed for sample collection using a protocol

approved by the Institutional Review Board of Penn State Hershey Cancer

Institute. Buffy coats from normal donors were also obtained from the blood

bank of the Milton S. Hershey Medical Center, Pennsylvania State University,

College of Medicine. Peripheral blood mononuclear cells (PBMCs) were isolated

by Ficoll-Hypaque gradient separation, as described previously [160]. Cell

41

viability was determined by trypan blue exclusion assay with more than 95%

viability in all the samples.

Cell culture

Freshly isolated PBMCs and primary CLL patient cells were cultured using RPMI-

1640 medium supplemented with 10% fetal bovine serum (both from Invitrogen).

JVM3 cells (DSMZ – German Collection of Microorganisms and Cell Cultures,

Braunschweig, Germany), a CLL cell line, were also cultured in this same

medium and cells were grown in 5% CO2 at 37˚C.

Preparation of nanoliposomal ceramide

12% pegylated nanoliposomes (80 ± 15 nm in size) that contain 30 mol%

ceramide were prepared as described previously with lipids 1,2-distearoyl-sn-

glycero-3-phosphocholine, 1,2-dioleoyl-sn-glycero-3-phosphoethanolamine, N-

hexanoyl-D-erythro-sphingosine (C6-ceramide), 1,2-distearoyl-sn-glycero-3-

phosphoethanolamine-N-[methoxy polyethylene glycol-2000], and N-octanoyl-

sphingosine-1-[succinyl(methoxy polyethylene glycol-750)] (PEG(750)-C8)

combined in chloroform at a molar ratio of 3.75:1.75:3:0.75:0.75 [159]. Combined

lipids were dried under nitrogen gas and resuspended in 0.9% sterile NaCl at

60°C. Following rehydration, resulting solution was sonicated for 5 min followed

by extrusion through a 100-nm polycarbonate membrane using the Avanti Mini

Extruder (Avanti Polar Lipids). Ghost liposomes were prepared in a similar

manner excluding N-hexanoyl-D-erythro-sphingosine (C6). Dihydro-C6-ceramide

liposomes were prepared in a similar manner by replacing N-hexanoyl-D-erythro-

42

sphingosine with N-hexanoyl-D-erythro-sphinganine. Several QA/QC parameters

were evaluated after preparation of nanoliposomes. Nanoliposomes were

formulated within the size range of 85nM – 90nM as measured by dynamic light

scattering. Zeta potentials of the nanoliposomes were measured and these were

between -3 mV to -7 mV, ensuring a neutral charge on the nanoliposomes.

Encapsulation efficiency of the ceramide nanoliposomes was evaluated by

LC/MS/MS.

Cell viability assay

Cell viability was performed using CellTiter 96® Aqueous One Solution assay kit

(Promega) or alamarBlue® assay kit (Invitrogen). Relative viable cell number

was determined by reading the plates at 490 nm or 570 nm wavelength

respectively in Synergy HT Multi-Detection Microplate Reader (Bio-TEK). All

samples were assayed in triplicate and each experiment was repeated at least

three times.

Apoptosis and exclusion assays (Annexin V/7AAD and TUNEL)

Apoptosis was determined in JVM3 cells and in PBMC samples from normal

donors (n=3) by 2-color flow cytometry with annexin-V and 7-amino-actinomycin

D (BD Pharmingen Transduction Laboratories) staining using 5 x 105 cells per

sample. TUNEL assays were also performed using terminal deoxylnucleotidyl

transferase (TdT), biotinylated UTP, and fluorescein isothiocyanate-streptavidin

as described previously [161]. Cell viability was also confirmed by trypan blue

exclusion.

43

Caspase 3/7 assay

For detection of caspase-3 and caspase-7 activation, JVM3 cells were plated in

replicates of six in 96-well plates, and treated with varying doses of ceramide for

24 hours and analyzed using ApoONE Homogeneous Caspase 3/7 Assay

(Promega) according to the manufacturer's instructions.

Western blot analysis

Western blot analyses of caspase 3, PARP, GAPDH, LDH, GLUT1, α-tubulin and

-actin protein expression were performed on whole-cell lysates collected using

RIPA buffer (Sigma). Densitometry analysis was performed using ImageJ

software.

Phase-contrast microscopy

To visualize a morphological necrotic phenotype, JVM3 cells were plated at 1 x

106 cells/ well and treated with 25 µM ghost nanoliposomes for 24 hours or C6-

ceramide nanoliposomes for 2, 6 and 24 hours. Phase contrast microscopy

images were then taken (Olympus CKX41).

GAPDH gene expression: real-time quantitative RT-PCR

Real-time reverse-transcription polymerase chain reaction (RT-PCR) was

performed using primer sets specific for GAPDH and an internal standard, 18S

rRNA, in an ABI PRISM 7900 sequence detector (Applied Biosystems) as

described elsewhere [144]. TRIzoL LS Reagent (Invitrogen) was used for RNA

extraction. Amplification of triplicate cDNA template samples was performed with

44

denaturation for 15 minutes at 95˚C, followed by 45 PCR cycles of denaturation

at 94˚C for 15 seconds, annealing at 55˚C for 30 seconds, and extension at 72˚C

for 30 seconds. A standard curve of cycle thresholds using serial dilutions of

cDNA samples was established and used to calculate the relative abundance of

the target gene. Values were normalized to the relative amount of 18S mRNA.

The relative amount of PCR products generated from each primer set was

determined based on the threshold cycle or threshold cycle value [162]. The

following primers were used for detection: GAPDH sense 5’-

GACCCCTTCATTGACCT CAACTACATG -3’, GAPDH antisense 5’-

GTCCACCACCCTGTTGCTGTAGCC-3’.

Glucose uptake and lactate production

The [3H]-2-DG uptake assay was modified from Kueck et al. [163]. 2-[1,2,-3H

(N)]-Deoxy-D-glucose was purchased from PerkinElmer. Cells were plated in 24-

well plates and 24 hours after experimental treatment were washed with PBS

and resuspended in Krebs-Ringer phosphate buffer + 1% BSA. Glucose uptake

was initiated by adding 1 µCi/mL 3H-2DG and 10 mM unlabeled glucose for 15

minutes. Uptake was terminated by adding ice cold Krebs-Ringer phosphate

buffer + 0.2 mM phloretin. 20 µM cytochalasin B, a potent inhibitor of glucose

transport was used as a control. All experiments were done in triplicate. Lactate

concentrations in culture media were determined after nanoliposomal treatment

using a colorimetric assay kit (Biomedical Research Service Center, SUNY at

Buffalo).

45

Measurement of ATP

Cellular ATP content was measured using a luminescence assay (Cell-Titer Glo

Kit, Promega). JVM3 cells were pretreated in the absence or presence of

pyruvate for 2 hours, and then incubated for 24 hours with nanoliposomal

ceramide. Final luminescence was measured in Synergy HT Multi-Detection

Microplate Reader (Bio-TEK).

shRNA Knockdown of GAPDH

GAPDH shRNA lentiviral clones (Human pLKO.1 vector) were purchased from

Open Biosystems (Huntsville, AL) and used to infect JVM3 cells according to the

manufacturer’s protocol. After selection with puromycin, a pool of infected cells

were expanded and tested for GAPDH expression via Western blotting.

Scrambled shRNA was used as a control in these experiments.

Lentiviral overexpression of GAPDH

Human pLOC vector (Open Biosystems) was used for overexpressing GAPDH in

JVM3 cells. Briefly, viral particles were produced in HEK293-FT cells using

pLOC, VSVG, tat and DR8.2 plasmids and JVM3 cells were transduced thrice

with the viral media according to the manufacturer’s protocol. JVM3 cells were

grown for 72 hours after the last transduction and subsequently harvested for

experiments. GAPDH overexpression was detected 72 hours after last

transduction via Western blotting. pLOC vector containing a red fluorescent

protein (RFP) sequence instead of GAPDH gene was used as a control in these

expreiments.

46

Animal Studies

Animal experimentation was performed according to protocols approved by the

Institutional Animal Care and Use Committee at the Pennsylvania State College

of Medicine and all efforts were made to minimize animal suffering. Female

Balb/c Nu/nu mice of about six weeks of age were obtained from Charles River

Laboratory (Wilmington, MA). Mice were subject to irradiation (600 cGy) one day

prior to inoculation. Ten million JVM3 cells were subcutaneously injected into the

right flank of the mice and treatment began approximately two weeks after

inoculation when tumors reached a volume of 50 to 100 mm3. Based upon prior

studies with multiple animal models [110, 144, 147, 159], leukemic mice were

treated with 40 mg/kg ghost (n=8) or C6-ceramide nanoliposomes (n=8) via IV

injection every other day over a three week period of time. Mice with large or

ulcerated tumors were euthanized.

In a parallel study, leukemic mice treated with an identical dose regimen of ghost

or C6-ceramide nanoliposomes were sacrificed during the course of the study

and tumor tissue was then flash frozen and isolated for protein extraction. Mice

were sacrificed on day 8 of C6-ceramide treatment (n=5), day 14 of C6-ceramide

treatment (n=6) and day 17 of C6-ceramide treatment (n=6). Additionally, mice

treated with the ghost nanoliposomes for 17 days were also sacrificed (n=5) and

protein was extracted from the tumor tissue for further immunoblot analysis.

47

Ethics Statement

A protocol approved by the Institutional Review Board of Penn State Hershey

Cancer Institute (Protocol #29839) was used to collect peripheral blood

specimens from CLL patients. Animal experiments were performed according to

a protocol (Protocol #2009-017) approved by the Institutional Animal Care and

Use Committee at the Pennsylvania State College of Medicine and all efforts

were made to minimize animal pain and discomfort.

Statistical analysis

All data are expressed as mean +/- SEM. All the graphs represent at least three

independent experiments, each replicated in triplicate, unless specified

otherwise. Paired Student t test (2-tail paired) and 2 way analysis of variance test

were used to determine the statistical significance and P value of </= 0.05 was

considered statistically significant.

48

Results

Nanoliposomal C6-ceramide selectively induces cell death in CLL cells

We have previously demonstrated the therapeutic use of C6-ceramide

nanoliposomes in both solid and non-solid tumor models [110, 144, 147, 159]. In

the present study, we investigated the therapeutic efficacy of this nanoliposomal

formulation in CLL. The CLL in vitro JVM3 cell line utilized for these studies was

established by EBV-transformation of human primary B-prolymphocytic leukemic

cells and treatment with phorbol ester, TPA [164]. MTT assay (Fig. 2-1A) and

trypan blue staining (Fig. 2-1B) demonstrated that nanoliposomal C6-ceramide,

but not the ghost nanoliposomes, induced dose-dependent cell death in JVM3

cells. Dihydro-C6-ceramide, a less active analog of C6-ceramide modestly

reduced cell viability at higher doses (Fig. 2-1A); however C6-ceramide was

significantly more toxic to cells in comparison. Concentrations of C6-ceramide

nanoliposomes above 25µM significantly increased percentage of non-viable

cells as shown by annexin V/7AAD staining (Fig. 2-1C) and TUNEL flow

cytometry analysis (Fig. 2-1D). It was further demonstrated that this dose-

dependent cell death induced by nanoliposomal C6-ceramide was also observed

in primary CLL patient cells but not in PBMC isolated from normal donors (Fig. 2-

1E). Flow cytometry analysis confirmed that apoptosis was not induced in these

normal cells (Fig. 2-1F). These results indicate that nanoliposomal C6-ceramide

is preferentially targeting CLL cells and is non-toxic to normal donor cells.

49

Figure 2-1. Nanoliposomal C6-ceramide selectively induces cell death in CLL cells. JVM3

cells were treated with varying doses of ghost or C6-ceramide or dihydro-C6-ceramide

nanoliposomes for 24 hours then A). MTT assay, B). Trypan blue staining was performed.

ANOVA statistical test was used to determine dose dependency between various C6-ceramide

treatment groups. P < .0001. C). JVM3 cells were treated with different doses of ghost or C6-

50

ceramide nanoliposomes for 24 hours, then cells were stained with annexin V and 7AAD for

apoptosis assay. D). Percentage of apoptotic cells was determined via TUNEL analysis after 24

hours. E). PBMC isolated from either CLL patients (n=3) or normal donors (n=3) were treated with

25 µM ghost or C6-ceramide nanoliposome for 2, 16 and 24 hours, then MTT assay was

performed. F). Percentage of apoptotic cells was determined in PBMC from normal donors (n=3)

via annexin V/7AAD staining. Cells were treated with different doses of ghost or C6-ceramide

nanoliposome for 24 hours or treated with 25 µM ghost or C6-ceramide nanoliposome for 2 and

24 hours. * P < 0.05.

Cell death induced by nanoliposomal C6-ceramide is independent of

caspase 3/7

To further characterize the mechanism of cell death being induced in the

JVM3 cell line, we investigated the effect of nanoliposomal C6-ceramide

treatment on caspase cleavage. Caspase 3 or downstream PARP cleavage was

not observed following treatment with the C6-ceramide nanoliposomes (Fig. 2-

2A). Treatment of JVM3 cells with 5 and 10 µM dasatinib, an inducer of apoptosis

in CLL [165], increased cleavage of both caspase 3 and PARP (Fig. 2-2B), thus

confirming that there was no defect in the caspase 3/7 apoptotic pathway in this

cell line. No change in caspase activity was also confirmed by the caspase 3/7

assay (Fig. 2-2C). As controls, 5µM dasatanib was sufficient to stimulate caspase

activity and pretreatment with 15µM zVAD-fmk, a pan-caspase inhibitor, inhibited

this activation. We further confirmed that C6-ceramide-induced cell death was

independent of caspase activation, as pre-treatment with zVAD-fmk did not

rescue cell death (Fig. 2-2D).

We next investigated if this caspase 3/7-independent cell death results in

a necrotic phenotype. To determine the predominant cell death mechanism being

51

induced following nanoliposomal ceramide treatment, JVM3 cells were treated

with 25 µM ghost or C6-ceramide nanoliposomes for varying time points and

visualized using phase contrast microscopy (Fig. 2-2E). Arrows indicate cells

which have increased in cellular volume and resemble a necrotic morphology as

evidenced by early plasma membrane rupture and abundance of cellular debris.

In addition, we confirmed necrotic cell death by flow cytometric analysis and

observed a time-dependent increase in Annexin V positive and 7AAD positive

cell population after treatment with nanoliposomal ceramide for 24 hours (data

not shown). From these results, we conclude that the preferential selectivity of

ceramide treatment appears to target a necrotic cell death mechanism,

consistent with caspase 3/7-independent cell death.

52

Figure 2-2. Cell death induced by nanoliposomal C6-ceramide occurs through caspase 3/7-

independent necrosis. JVM3 cells were treated with A). different doses of ghost and C6-

ceramide nanoliposome for 24 hours, B) 5 µM and 10 µM dasatinib, as well as DMSO vehicle

control for 24 hours, then Western Blot analysis was performed for caspase 3 and PARP

cleavage. C). Enzymatic activities of caspase 3/7 were measured using caspase 3/7

luminescence kit. The results for cells treated with zVAD-fmk and zVAD-fmk + Dasatinib were

significantly different from untreated cells (NT) and cells treated with Dasatinib. * P < 0.05. D).

JVM3 cells were treated with varying doses of ghost or C6-ceramide nanoliposomes for 24 hours

following a 2 hour pre-treatment or no pre-treatment with zVAD-fmk (15 µM). Cell viability was

assessed via MTT assay. The viability of cells treated with ceramide liposomes and cells treated

53

with ceramide liposomes +zVAD-fmk was significantly different from viability of cells treated with

ghost liposomes * P < 0.05. E). JVM3 cells were treated with 25 µM ghost nanoliposomes for 24

hours or C6-ceramide nanoliposomes for 2, 6 and 24 hours and phase contrast microscopy

images were taken. Arrows indicate morphology of necrotic cell death.

GAPDH as a target for treatment of CLL

There is growing evidence suggesting a link between increased

dependency on glucose metabolism and overexpression of glycolytic enzymes in

cancer cells. Moreover, GAPDH has been shown to be upregulated in many

cancers and is the primary target of some chemotherapeutic drugs [166].

Recently, GAPDH has been implicated in promoting cellular survival,

chemotherapy-resistance and protection from caspase-independent cell death

[167, 168]. We now demonstrate that treatment of JVM3 cells with nanoliposomal

C6-ceramide leads to a significant decrease in GAPDH protein expression (Fig.

2-3A). There was a significant decrease in GAPDH protein following 8 hours

treatment with 25µM nanoliposomal C6-ceramide (Fig. 2-3A(i)).Treatment with

varying concentrations of nanoliposomal C6-ceramide for 24 hours significantly

decreased GAPDH protein expression (Fig. 2-3A(ii)). Based upon these

preliminary studies, we chose to evaluate the physiological and therapeutic

consequences of C6-ceramide-mediated GAPDH reduction at 24 hours as a

function of concentration. It was also apparent that GAPDH protein expression

was decreased following treatment with C6-ceramide nanoliposomes in primary

CLL cells (Fig. 2-3B). In addition, using qRT-PCR analysis it was determined that

mRNA expression of GAPDH in JVM3 cells was decreased after treatment with

nanoliposomal C6-ceramide (Fig. 2-3C). Treatment with nanoliposomal C6-

54

ceramide did not decrease GAPDH protein expression in non-transformed cells

from normal donors (Fig. 2-3D), which again alludes to the preferential specificity

of nanoliposomal C6-ceramide treatment for cancer cells. Reduction in GAPDH

protein preceded induction of cell death. GAPDH reduction was observed as

early as 8 hours after treatment with nanoliposomal C6-ceramide, whereas C6-

ceramide nanoliposomes induced cell death 12 hours after treatment (data not

shown). This indicated that C6-ceramide-induced decrease in GAPDH played a

critical role to induce cell death

A variety of glycolytic genes are ubiquitously overexpressed in human

cancers and are thought to contribute to enhancement of glycolysis [151]. The

differential expression of glycolytic enzymes in CLL was previously unknown, so

we sought to investigate the expression of GAPDH in CLL. The basal protein

level of GAPDH was compared between normal donor PBMC (n=8) and PBMC

isolated from CLL patients (n=14). CLL patients were further stratified according

to their white blood count (WBC) count at the time their plasma was collected.

Patients with WBC counts >50,000 cells/µL were categorized as having high

WBC count (n=5), while patients with counts <50,000 cells/µL were in the low

WBC count group (n=9). It was evident that GAPDH protein expression was

significantly overexpressed in the CLL patient population with high WBC count

(Fig 2-3E). Moreover, treatment with 25µM C6-ceramide nanoliposomes was

significantly more cytotoxic in the subset of CLL patients with the higher WBC

count (Fig 2-3F). Collectively, our data suggest that nanoliposomal C6-ceramide

55

is targeting GAPDH and presents a new mechanism for which ceramide is

selectively targeting cancer cells.

Figure 2-3. C6-ceramide nanoliposomes target GAPDH in CLL. A). JVM3 cells were treated

with (i) 25µM of ghost or C6-ceramide nanoliposomes for varying times, (ii) varying doses of

56

ghost or C6-ceramide nanoliposomes for 24 hours, then Western Blot analysis was performed for

GAPDH. Densitometry analysis from replicate experiments (n=4) was performed via ImageJ

software. In Fig. A(ii) the representative blot demonstrates a decrease in GAPDH at 12.5µM C6-

ceramide, a result not observed in the other two distinct replicate experiments. The protein levels

are also quantified in the graphs. * P < 0.05; ** P <0 .005. B). CLL patient cells (n=3, analyzed

individually) were treated with different doses of ghost or C6-ceramide nanoliposomes for 24

hours, then Western Blot analysis was performed for GAPDH. * P < 0.05. The blot represents the

effect of C6-ceramide nanoliposomes on one patient sample, however, the graph is an average of

the effect on all three patient cells. C). JVM3 cells were treated with different doses of ghost and

C6-ceramide nanoliposomes for 24 hours, then qRT-PCR analysis was performed for expression

of GAPDH mRNA. Expression was normalized to 18S. * P < 0.05. D). PBMC from normal donors

(n=3) were treated with 25 µM ghost or C6-ceramide nanoliposome for 24 hours and then

Western Blot analysis was performed for GAPDH. E). Basal protein expression of GAPDH was

determined via Western Blot analysis on PBMC isolated from normal donors (n=8) or from CLL

patients with either a lower WBC count (n=9) or higher WBC count (n=5). Patients with WBC

counts >50,000 cells/µL were identified as patients with high WBC count. * P < 0.05. F). CLL

patient cells with either a lower WBC count (n=4) or higher WBC count (n=3) were treated with

25µM ghost or C6-ceramide nanoliposomes for 2, 16 and 24 hours then an MTT assay was

performed * P < 0.05.

C6-ceramide targets GAPDH-dependent glycolysis

To further demonstrate that C6-ceramide was targeting the glycolytic

pathway, consistent with our observation of a decrease in GAPDH expression,

we investigated the effects of C6-ceramide upon lactate and ATP production.

Treatment with varying doses of nanoliposomal C6-ceramide for 24 hours

caused a significant decrease in lactate concentration (Fig. 2-4A) and a dose-

dependent decrease in ATP production (Fig. 2-4B) in JVM3 cells. 3-

bromopyruvate, a glycolytic inhibitor was used as a positive control in the ATP

production experiment. To demonstrate that decreased ATP production was

dependent upon ceramide-reduced GAPDH expression we utilized a lentiviral

57

approach to decrease GAPDH in JVM3 cells (Fig. 2-4C inset). We observed a

comparable decrease in ATP production between JVM3 cells treated with

GAPDH shRNA and 25 µM C6-ceramide nanoliposomal treated JVM3 cells. C6-

ceramide nanoliposomal treatment further reduced ATP production in GAPDH-

silenced JVM3 cells (Fig. 2-4C). Taken together, these results suggest that

ceramide decreased ATP production and it may be partially GAPDH-dependent.

Increasing levels of intracellular ceramide correlates with downregulation

of nutrient transporters and leads to cellular starvation [148, 163]. We wanted to

determine if ceramide was also targeting glucose transport in addition to

targeting the GAPDH glycolytic enzyme. Results from a tritiated glucose uptake

assay showed no significant difference between JVM3 cells treated with ghost or

C6-ceramide nanoliposomes (Fig. 2-4D). A significant decrease was observed,

however, in cells treated with cytochalasin B, a potent inhibitor of glucose

transport. Moreover, protein expression of the glucose transporter, GLUT1, was

not altered after treatment with nanoliposomal C6-ceramide (Fig 2-4E). Further

investigation into other glycolytic enzymes, like lactate dehydrogenase (LDH) and

pyruvate kinase M2 (PKM2), revealed no difference between the ghost- and C6-

ceramide nanoliposome treated groups (data not shown).

To illustrate that cell death induced by nanoliposomal C6-ceramide was

dependent on targeting of the glycolytic pathway, we pre-treated JVM3 cells with

pyruvate, the ultimate downstream product of glycolysis and determined its effect

on cell viability and ATP production. Pre-treatment of cells with 10 mM pyruvate

led to a significant rescue of cell viability (Fig. 2-5A) and ATP production (Fig. 2-

58

5B) after C6-ceramide nanoliposome treatment. To further demonstrate that

reduction in GAPDH protein played a critical role in C6-ceramide-induced cell

death, we overexpressed GAPDH in JVM3 cells using a lentiviral transduction

overexpression system (Fig. 2-5C). We showed by Western blotting that

transduction with lentiviral particles containing the GAPDH gene (Lenti-GAPDH)

resulted in overexpression of GAPDH protein (Fig. 2-5C insert). Surprisingly, the

procedure itself (control lentiviral particles containing RFP gene, Lenti-RFP) also

elevated GAPDH protein expression (Fig. 2-5C insert). In both cases,

nanoliposomal C6-ceramide was less effective in inducing cell death when

GAPDH protein was elevated. Taken together, overexpression of GAPDH

significantly rescued cell death in JVM3 cells after treatment with nanoliposomal

C6-ceramide (Fig. 2-5C).

59

Figure 2-4. C6-ceramide targets the glycolytic pathway. JVM3 cells were treated with varying

doses of ghost or C6-ceramide nanoliposomes for 24 hours, then A) lactate production was

analyzed, B) ATP production was analyzed; * P < 0.05. The graphs depict average results from

three independent experiments. Equal numbers of cells were seeded in each well during the

experiments for the purpose of normalization. C). GAPDH was effectively knocked down in JVM3

cells via a lentiviral shRNA approach (inset). ATP production was then assessed after 24 hours of

treatment with 50µM of ghost or 25µM of C6-ceramide nanoliposomes; * P < 0.05; ** P < 0.005.

D). Glucose uptake was assessed after JVM3 cells were treated with ghost and C6-ceramide

nanoliposomes for 24 hours. Cytochalasin B was used as a positive control. E). JVM3 cells were

treated with ghost or C6-ceramide nanoliposomes for 24 hours, then Western Blot analysis was

performed for GLUT1.

60

Figure 2-5. Pharmacological and molecular confirmation that nanoliposomal C6-ceramide

targets the glycolytic pathway at the level of GAPDH. A). JVM3 cells were pre-treated for 2

hours with 10mM pyruvate, then treated with ghost or C6-ceramide nanoliposomes for 24 hours.

MTT cell viability assay was performed. B). ATP production in these cells was also determined;

61

The results are significant between cells treated with ceramide liposomes and cells treated with

ceramide and pyruvate, ** P < 0.005. C). JVM3 cells were transduced with lentiviral particles

overexpressing GAPDH or viral particles expressing RFP (control). Post transduction, Western

blotting analysis was done to determine levels of GAPDH in experimental and control cells

(insert). Cells were treated with ghost or C6-ceramide nanoliposomes for 24 hours. Cell viability

was determined using MTT assay and AlamarBlue assay; * P < 0.005.

Nanoliposomal C6-ceramide displays anti-leukemic effect in CLL animal

model

Given that we have demonstrated selective induction of cell death utilizing

nanoliposomal ceramide in CLL cells in vitro, we next investigated this

therapeutic approach in an in vivo model. Moreover, this model would allow us to

confirm the selective actions of ceramide upon GAPDH-dependent glycolysis in

vivo. Loisel et al. have reported the establishment of a novel human CLL like

xenograft model in nude mice using JVM3 cells [169]. We evaluated the ability of

C6-ceramide nanoliposomes to inhibit tumor growth of leukemic CLL cells in

xenografts in this immunodeficient mouse model. C6-ceramide nanoliposomes

decreased tumor growth beginning on day 13 of treatment (Fig. 2-6A). In

addition, C6-ceramide treated tumors did not progress for the remainder of the

study. Immunoblot analysis showed that C6-ceramide, but not ghost

nanoliposomes, significantly reduced GAPDH protein expression at day 8, 14

and 17 (Fig. 2-6B). ImageJ densitometry analysis for multiple animals is shown,

as well as a representative blot from Day 17 (Fig. 2-6B inset).

Collectively, these results indicate that bioactive ceramide analogues can

be incorporated into pegylated nanoliposomal formulations and elicit potent anti-

62

leukemic efficacy in a mouse model of CLL. In addition, these results

substantiate our previous in vitro data that cell death induction after treatment

with C6-ceramide nanoliposomes involves targeting of GAPDH and the glycolytic

pathway in CLL.

Figure 2-6. Nanoliposomal C6-ceramide displays anti-leukemic effect in a CLL animal

model. A). Two weeks after ten million JVM3 cells were inoculated in the right flank of female

Balb/c Nu/nu mice, animals were treated with 40 mg/kg ghost (n=8) or C6-ceramide (n=8)

63

nanoliposomes via tail vein injection. Treatment regimen was every other day over a three week

period of time. Tumor size was assessed every other day; * P < 0.05; ** P < 0.005. B). Leukemic

mice following an identical dose regimen were sacrificed on day 8 (n=5), day 14 (n=6) and day 17

(n=6) of C6-ceramide treatment and on day 17 for ghost treatment. Immunoblot analysis for

GAPDH protein expression was performed on tumor tissues. A representative blot from day 17 is

shown (Fig. 6B inset); * P < 0.05.

Discussion

The present study identifies dysregulated glucose metabolism as a novel

target in CLL. Furthermore, we found that treatment with C6-ceramide

nanoliposomes led to preferential induction of caspase-independent, necrotic cell

death in CLL cells in vitro. Several investigations support a caspase-

independent, necroptotic cell death mechanism for chemotherapeutics

(tamoxifen, fludarabine) that generate endogenous ceramide [137, 170]. In

addition, short chain ceramides induce necrotic cell death independent of

caspase 3 activation in B cell lymphomas [36]. Several groups, including ours

have studied the metabolism of exogenous short chain ceramides in cancer cells.

Exogenous C6-ceramide is metabolized into natural ceramides through de-

acylation to yield sphingosine followed by subsequent re-acylation with various

fatty acids [171]. Exogenous C6-ceramide is also metabolized to C6-

sphingomyelin, C6-glucosylcreamide and cerebrosides [112, 171, 172]. The

predominant metabolic pathway of exogenous C6-ceramide is specific to the

cancer cell type and concentration of C6-ceramide delivered to the cells.

We observed that treatment of CLL cells with nanoliposomal C6-ceramide

decreased protein and mRNA expression of GAPDH and based upon initial

64

experiments (Fig. 3A) we chose to evaluate the physiological consequences of

ceramide-mediated decrease in GAPDH at 24 hours as a function of various

concentrations of ceramide. GAPDH is known to mediate glycolysis and is

responsible for oxidizing glyceraldehyde-3-phosphate to 1,3-biphosphoglycerate.

Until recently, GAPDH has been considered as a constitutive housekeeping gene

and was used as a control for normalizing changes in expression of protein and

genes. Current studies suggest that GAPDH is upregulated in many cancers and

is the primary target of some chemotherapeutic drugs [166]. The tumor

suppressor TP53 has been shown to increase GAPDH expression in endothelial

cells [173] and significant increases in GAPDH expression were also observed in

breast cancer cells stimulated with growth factors [174]. Interestingly, GAPDH

plays a key role in opposing caspase-independent cell death and promoting

cellular survival [167]. It has been suggested that GAPDH overexpression may

assist cancer cells in evading apoptosis and cell death mechanisms [166]. This

protection from cell death is mediated by an elevation in glycolysis and enhanced

ATP levels [168]. Moreover, it has been demonstrated that overexpression of

GAPDH is sufficient to protect chronic myelogenous leukemia (CML) cells from

imatinib-induced caspase-independent cell death and knockdown of GAPDH

resensitizes cells to imatinib-induced cell death [175]. Therefore, combining a

drug that reduces GAPDH expression (like ceramide) with imatinib could prove to

have therapeutic benefit in overcoming this drug resistance.

In addition to its role in the glycolytic pathway, GAPDH has also been

shown to serve other functions in the cell which remain to be elucidated.

65

Although generally in a cytosolic form, the GAPDH protein can also be found in

the nucleus and nuclear functions of GAPDH include transcriptional regulation,

DNA repair, and maintenance of telomere structure [176]. There is rapid

shortening of telomere length when cells are treated with exogenous, short-chain

ceramides [177]. Ogretmen and colleagues demonstrated that overexpression of

GAPDH results in protection of telomeric DNA in response to exogenous

ceramide treatment [177]. Further analysis showed that ceramide was inhibiting

nuclear localization of GAPDH and showed that a potential mechanism of

ceramide-mediated shortening of telomeres in somatic cells leads to cell

senescence while maintenance of telomeres is associated with immortality of

cancer cells. These data suggest another mechanism by which increased

GAPDH expression contributes to a survival advantage in cancer cells.

Elevation of glucose uptake and glycolysis in cancer cells can depend

heavily on the upregulation of glucose transporters [178]. Others have shown

that nutrient transporter down-regulation is critical in ceramide-induced cell death

[148], so we sought to clarify this hypothesis in CLL. Using a radiolabeled

glucose uptake assay, we demonstrated that the ability of ceramide to induce cell

death in CLL was not due to blocking nutrient transport of glucose into the cell. In

addition, expression of the GLUT1 glucose receptor was not regulated by

ceramide. This lack of effect upon nutrient transport further suggests direct

effects upon glycolytic enzymes overexpressed in cancer.

We identified GAPDH as being overexpressed in a population of CLL

patients. In addition to the Western blot analysis, we also re-analyzed previous

66

microarray data from CLL patients [179] using Gene Expression Omnibus (GEO)

to investigate dysregulation of glycolytic enzymes in CLL. Interestingly, we found

that all of the enzymes in the glycolytic pathway were upregulated in the CLL

patients with a poor prognosis (data not shown). Based on our results and the

independent microarray data, we conclude that glycolysis appears to be

upregulated in a subset of CLL patients with either higher white blood cell counts

and/or a more aggressive clinical course.

Our findings demonstrate that the cytotoxicity of C6-ceramide is due to

targeting of glucose metabolism in CLL. Previous studies reported that inhibitors

of ATP production, like bromopyruvate, effectively induce cell death in cancer

[180]. The abolishment of cellular ATP production proves effective due to the

increased dependence on glycolysis and ATP production in oncogenic cells.

Moreover, treatment with koningic acid, an inhibitor of GAPDH selectively

induces cell death in cancer via ATP deprivation [181]. In a similar fashion, we

also show that nanoliposomal C6-ceramide treatment induces cell death in CLL

through targeting of the glycolytic pathway and results in decreased ATP and

lactate production. Even though our studies indicate that shRNA knock down of

GAPDH reduced ATP production similar to 25 µM C6-ceramide nanoliposomal

treatment, the fact that ceramide augmented this shRNA approach suggests that

ceramide may be targeting the Warburg effect through both decreased GAPDH

as well as an undefined secondary mechanism.

To better understand and confirm if inhibition of the glycolytic pathway was

the means by which ceramide was inducing cell death in CLL, we carried out

67

experiments that involved pre-treatment with a final end product of glycolysis. It

was determined that pre-treatment of CLL cells with pyruvate was sufficient to

rescue ATP depletion and cell death that was otherwise induced after treatment

with the C6-ceramide nanoliposome. In addition, we transiently overexpressed

GAPDH in CLL cells to demonstrate rescue of C6-ceramide-induced cell death in

these cells. Interestingly, we also observed rescue of cell death in control cells

transduced with control viral particles. Yet, protein analysis showed that both, the

experimental and control cells overexpressed GAPDH. This alludes to the fact

that transient overexpression of GAPDH partially protects CLL cells from C6-

ceramide-induced cell death.

Confirming in vitro studies, we provide clear evidence that

nanoliposomal C6-ceramide demonstrated in vivo efficacy via tumor growth

inhibition in a murine xenograft model of CLL. We are aware of a recently

described mouse model for CLL generated by co-injection of primary human CLL

cells and T cells into NOD/ SCID mice, but have chosen to focus our studies on

the mouse xenograft model because we can obtain tissue for ex vivo Western

analysis of GAPDH protein levels [182]. In our in vivo studies we observed that

treatment with nanoliposomal C6-ceramide effectively decreased tumor burden

without systemic side effects. This is consistent with previous studies from our

laboratory and others which show that this C6-ceramide nanoliposome

formulation results in tumor regression in animal models of cancer and

hematological malignancies [144, 147, 159]. In addition, it was expected that this

nanoliposomal formulation would be relatively non-toxic to the animals, as it was

68

previously selected by the Nanotechnology Characterization Laboratory (National

Cancer Institute) for extensive toxicology and stability testing. (Detailed

information on the toxicology studies of the “Ceramide Liposomes” can be found

at http://ncl.cancer.gov/working_technical_reports.asp). In vivo studies also

confirm that ceramide is targeting GAPDH in CLL, as protein isolated from tumor

tissue showed an overall decrease in GAPDH expression following treatment

with C6-ceramide nanoliposomes.

In conclusion, we show C6-ceramide nanoliposomes preferentially inhibit

the altered metabolism of glucose in leukemic cells via downregulation of

GAPDH, resulting in induction of necrotic cell death. We provide the first

evidence that GAPDH is overexpressed in a subset of CLL patients. Our findings

provide a metabolic explanation for the increased sensitivity and selectivity of

cancer cells to ceramide. Taken together, these results suggest that

nanoliposomal C6-ceramide could be an effective novel therapy for patients

whose cancer cells overexpress GAPDH, including those with CLL.

69

CHAPTER 3: STAT3 mediates nanoliposomal C6-ceramide-induced cell death in chronic lymphocytic leukemia

Introduction

Chronic lymphocytic leukemia (CLL) is a slow growing B cell malignancy

and the most prevalent form of adult leukemia in the Western world [125]. It is

characterized by the clonal expansion and accumulation of mature B

lymphocytes in the bone marrow, peripheral blood and often the lymph nodes

[136]. Leukemic B cells express CD5, CD19, CD20 and CD23 on the surface and

are characterized by cell-cycle arrest in the G0/G1 phase and resistance to

programmed cell death [136, 183]. Depending on the degree of somatic

hypermutation and chromosomal abnormalities, the clinical course of CLL ranges

from an indolent and slow progressing disease to rapid disease progression [125,

136]. The standard treatment regimen for CLL includes a combination chemo-

immunotherapy with fludarabine, cyclophosphamide and rituximab with an overall

response rate of approximately 90% and complete remission of 72% [132, 133].

Despite these advances in therapeutics, eventual drug resistance and relapse

ultimately cause CLL to be an incurable and chronic disease with a need for

novel therapies [136].

Sphingolipids, a class of complex cellular lipids, function as structural

components in cellular membranes and as signal transducers regulating cellular

processes like growth, proliferation, response to stress, differentiation,

senescence, autophagy and apoptosis [2, 12]. Ceramide has been termed as a

“tumor-suppressor sphingolipid” because of its ability to potentiate signaling

70

cascades that lead to cell death [184]. Sphingolipid-based cancer therapeutics

are thus gaining increasing attention due to their critical role in cancer initiation,

progression, pathogenesis and metastasis. Sphingolipid-based cancer therapies

aim to perturb the native sphingolipid balance towards anti-proliferative and

apoptotic signaling by elevating levels of intracellular ceramide [18]. Our

laboratory has demonstrated that a nanoliposomal formulation of short chain C6-

ceramide (CNL) is an effective anti-tumorigenic agent in vitro and leads to tumor

regression in vivo in several animal models of cancer [35, 39, 110, 111, 147, 159,

185]. Specifically, we have demonstrated that CNL selectively targets the

Warburg effect in CLL cells by causing downregulation of glyceraldehyde 3-

phosphate dehydrogenase and elicits tumor regression in an in vivo murine

model of CLL [39]. There is also evidence that another short chain ceramide

analog, C2-ceramide, induces non-apoptotic cell death in CLL cells [137].

Interestingly, treatment with fludarabine, a chemotherapeutic agent often used as

first line of treatment for CLL also causes a 2.5 – 3 fold increase in intracellular

ceramide levels 6 hours after treatment and inhibiting accumulation of

intracellular ceramide prevents fludarabine-induced apoptosis [55, 137].

Altogether, these evidence suggest that accumulation of endogenous pro-

apoptotic ceramide species or delivery of pro-apoptotic exogenous ceramide is

induces cell death in CLL cells. Thus, CNL could potentially be an effective anti-

tumorigenic agent for CLL.

Signal transducer and activators of transcription (STAT) are latent

transcription factors that translocate to the nucleus upon activation by receptor or

71

non-receptor tyrosine kinases. The STAT signal transduction pathway plays a

critical role in hematopoietic biology because it mediates the response to a

multitude of cytokines [186]. Among the seven STAT members, STAT3 and

STAT5 have been studied extensively for their role in cancer progression [187].

Specifically in CLL, a report in 1997 demonstrated constitutive phosphorylation of

serine 727 (S727) but not tyrosine 705 (Y705) in STAT1 and STAT3 [188].

However, studies addressing the physiologic significance of this modification

have been conducted only in the last few years. It has been shown STAT3 that is

constitutively phosphorylated at S727 also has the ability to bind DNA and

activate transcription in CLL cells independent of Y705 phosphorylation [189]. In

line with the newly found function of STAT3 phosphorylated at S727 (p-STAT3-

S727) as a regulator of the respiratory chain in mitochondria, it has been

reported that p-STAT3-S727 in CLL cells associates with complex I of the

respiratory chain and imparts viability and stress protection to CLL cells [190,

191]. STAT3 also regulates micro-RNA expression in CLL cells [192].

Interestingly, unphosphorylated STAT3 also promotes CLL cell survival and

proliferation by activating the NF-ΚB pathway [193]. Additionally, ample evidence

in the literature demonstrates the critical role of STAT3 in mediating pro-survival

interactions between CLL cells and the microenvironment [194, 195]. STAT3

inhibitors have shown to sensitize CLL cells to apoptosis and also reverse IL6-

induced resistance to histone deacetylase inhibitors [196, 197]. Taken together, it

is clearly evident that the critical role of STAT3 in CLL pathogenesis makes it a

promising therapeutic target in this disease.

72

In this study, we sought to identify the molecular basis of CNL-induced cell

death in CLL. We demonstrate that CNL specifically suppresses phosphorylation

of STAT3 at S727 and Y705 residues, leading to a reduction in the transcriptional

activity of STAT3 and a subsequent downregulation of critical anti-apoptotic

proteins like Mcl-1 and survivin. Additionally, we have identified the upstream

kinases that are suppressed by CNL, eventually leading to reduction in STAT3

phosphorylation. CNL suppresses the activity of Bruton’s tyrosine kinase (BTK),

mitogen-activated protein kinase kinase (MEK) and protein kinase C (PKC)

resulting in a suppression of STAT3 phosphorylation. These findings collectively

indicate that CNL can potentially be an effective therapy for CLL. Our work

reports two novel targets of CNL. Firstly, these results demonstrate an effect of

CNL on BTK, a critical kinase mediating the B-cell receptor signaling in CLL cells.

Furthermore, the current use of ibrutinib, a BTK inhibitor, in the clinic for CLL

patients imparts clinical relevance to this finding. Secondly, we also demonstrate

an inhibitory effect of CNL on STAT3 phosphorylation.

Materials and Methods

Reagents

Antibodies specific for STAT3, p-STAT3-S727, p-STAT3-Y705, Mcl-1, Ran,

STAT1, p-STAT1-Y701, p-STAT1-S727, STAT2, p-STAT2-Y690, STAT5, Akt-

S473, BTK, p-BTK-Y223, p-ERK, ERK, p-MARCKS, MARCKS, survivin, XIAP,

cyclin D1, p21 and β-actin were purchased from Cell Signaling Technology Inc.

(Beverly, MA). For Western blotting, 2-12% or 12% precasted Nupage

electrophoresis gels were purchased from Invitrogen (Carlsbad, CA) and

73

chemiluminescence reagent was obtained from Thermo Scientific (Waltham,

MA). STAT3 inhibitor - Stattic, MEK inhibitor - U0126 and PKC inhibitor - Bis-I

were purchased from Sigma (St. Louis, MO). BTK inhibitor, ibrutinib was

purchased from MedChem Express (Monmouth Junction, NJ).

Patient characteristics and preparation of peripheral blood mononuclear

cells

All patients met the clinical criteria of CLL and were not on treatment at the time

of sample acquisition. Peripheral blood specimens from CLL patients were

obtained and informed consents signed for sample collection using a protocol

approved by the Institutional Review Board of Penn State Hershey Cancer

Institute. CLL PBMCs were chosen according to the following criteria: CD19+ >

80%, CD20+ > 80%, CD5+ > 90%. These criteria ensured that the PBMCs

isolated from CLL patient blood predominantly consisted of leukemic B-cells.

Buffy coats from normal donors were also obtained from the blood bank of the

Milton S. Hershey Medical Center, Pennsylvania State University, College of

Medicine. Peripheral blood mononuclear cells (PBMCs) were isolated by Ficoll-

Hypaque gradient separation, as described previously [160].

Cell culture

Culture of freshly isolated PBMCs and primary CLL patient cells was carried out

using RPMI-1640 medium supplemented with 10% fetal bovine serum (both from

Invitrogen). JVM-3 cells (DSMZ – German Collection of Microorganisms and Cell

Cultures, Braunschweig, Germany), a CLL cell line with wild type p53, were also

74

cultured in this same medium and cells were grown in 5% CO2 at 37˚C. Mec-2

cells (DSMZ – German Collection of Microorganisms and Cell Cultures,

Braunschweig, Germany), a CLL cell line with mutated p53 were cultured in

Iscove's MDM media supplemented with 10% FBS. HEK-293FT cells (Invitrogen,

Waltham, MA) were cultured in D-MEM media supplemented with 10% FBS and

1X Anti-anti antibiotic (Gibco, Waltham, MA).

Preparation of nanoliposomal ceramide

Preparation of nanoliposomal ceramide has been described in Chapter 2 of the

dissertation.

Preparation of lipid:BSA complexes

Lipids were dried under a stream of nitrogen and then resuspended in DMSO.

The lipid/DMSO solution was then added to a fatty acid-free BSA solution to

achieve a final concentration of 1 mM lipid, 1 mM BSA in 20 mM HEPES with

10% DMSO. Complexes were allowed to form by rocking at room temperature for

30 min, followed by sonication to clarity.

Cell viability assay

A set of experiments were conducted to determine the toxicity of the CNL and

Stattic in JVM-3 cells, Mec-2 cells, CLL patient cells and in normal donor PBMC.

Cell viability was performed using CellTiter 96® Aqueous One Solution assay kit

(Promega) and relative viable cell number was determined by reading the plates

at 490 nm wavelength in Synergy HT Multi-Detection Microplate Reader (Bio-

75

TEK). All samples were assayed in triplicate and each experiment was repeated

at least three times.

Cell death assays (Flow cytometry for AnnexinV/7AAD)

Apoptosis was determined in JVM-3 cells, Mec-2 cells and in CLL patient cells by

2-color flow cytometry with Annexin-V-PE and 7-amino-actinomycin D (BD

Pharmingen Transduction Laboratories) staining using 5 x 105 cells per sample.

Necrosis was quantidied in the Annexin V positive and 7AAD positive quadrant.

Data were collected by flow cytometry.

Western blot analysis

Western blot analysis was performed on whole-cell lysates collected using RIPA

buffer (Sigma). Blots were washed and developed with enhanced

chemiluminescence (Thermo Scientific, Waltham, MA) following the

manufacturer’s instructions. Densitometry analysis was performed using ImageJ

software.

shRNA Knockdown of STAT3

STAT3 shRNA plasmid clones (Human pTRIPZ vector) were purchased from

Open Biosystems (Huntsville, AL) and used to transfect JVM-3 cells.

Nucleofection was performed using the Amaxa Nucleofector I device. JVM-3

cells (3 × 106 ), resuspended in 100μl of Cell line Solution Kit V (Amaxa,

Cologne, Germany) with 6μg of shRNA were transfected with the Amaxa

Nucleofector I device (program X-001), cultured in six-well plates in complete

76

medium for 24, 48, 72 and 96 hours and then examined for STAT3 knockdown

using Western blot analysis. Cells were also analyzed for percent viability using

flow cytometric analysis for AnnexinV and 7AAD at 24, 48, 72 and 96 hours after

nucleofection.

Preparation of pervandate

1mM pervanadate stock was prepared by adding 10µL of 100mM sodium

orthovanate, 0.3% hydrogen peroxide diluted in 20mM HEPES and 940µL of

water. After 5 minutes of incubation, a small amount of catalase was mixed in the

pervanadate stock to remove excess hydrogen peroxide. The pervanadate was

used within 2 hours of preparation.

Luciferase reporter assay

Cignal reporter assay kit from Qiagen (Hilden, Germany) was used for obtaining

plasmids for the luciferase reporter assay. JVM-3 cells (2x106 cells) were

transfected with 4µg of either reporter construct, negative control construct or

positive control construct using the Amaxa Nucleofector I device (X-001

program). Cell were allowed to be in culture for 24 hours post transfection and

then treated for 12 hours with CNL or ghost liposomes. Dual-Glo luciferase assay

system from Promega (Madison, WI) was used to obtain luciferase

luminescence. The assay and quantification was done following the

manufacturer’s instructions.

77

Lentiviral transduction for STAT3C overexpression

Human EF.STAT3C.Ubc.GFP vector from Addgene (Cambridge, MA) was used

for expressing STAT3-C in JVM-3 cells [32]. Briefly, viral particles were produced

in HEK293-FT cells using pLOC, VSVG, tat and DR8.2 plasmids and JVM-3

cells were transduced thrice with the viral media according to the manufacturer’s

protocol. JVM-3 cells were grown for 72 hours after the last transduction. The

STAT3-C overexpression vector has an EGFP sequence as a selectable marker

and the transduced cells were sorted for EGFP and grown as a pure population

(JVM3-STAT3C cells). The cells were then harvested for experiments. Human

pLOC overexpression vector (Open Biosystems) containing a RFP sequence

was used as a negative control. Cells were treated with CNL or ghost liposomes.

JVM-3 cells were transduced thrice with media containing the negative control

virus. FACS was not performed for the control JVM3-RFP cells since we

obtained approximately 70-80% transduction efficiency. 72 hours after the last

transduction, JVM3-RFP cells were harvested for experiments. JVM3-STAT3C

cells and JVM3-RFP cells were treated with 20µM and 40µM CNL and ghost

liposomes for 24 hours. 24 hours after treatment, cells death was analyzed using

flow cytometric analysis using AnnexinV-V450 and 7AAD staining.

Statistical analysis

All data are expressed as mean plus or minus SEM. Paired Student t test (2-tail

paired) and 2-way analysis of variance test were used to determine the statistical

78

significance and P value of 0.05 or less was considered statistically significant.

All results are a mean of three independent biological triplicates unless specified.

Results

STAT3 is a potential therapeutic target in CLL

Several reports suggest that STAT3 might play a role in the pathogenesis of

CLL [189, 190, 197]. We compared levels of total STAT3 between normal cells

and CLL cells. STAT3 was highly overexpressed in both, CLL cell lines and

patient cells in comparison to peripheral blood mononuclear cells (PBMC)

obtained from normal blood donors (Fig. 3-1A). Next, we evaluated if inhibiting

STAT3 signaling in CLL cells would induce cell death. Knock down of STAT3 in

JVM3 cells by using an inducible lentiviral STAT3-shRNA significantly increased

the number of Annexin V positive cells 24 hours after doxycycline induction (Fig.

1B). An average of 57% knockdown of STAT3 protein was observed 24 hours

after induction (Fig. 3-1B). Doxycycline was non-toxic to JVM3 cells at doses

used for induction. STAT3 knockdown caused a 108% increase in cell death 24

hours after doxycycline induction as compared to control cells. As the

knockdown weakened at later time points, cell death induction reduced to 42%

48 hours after doxycycline induction, and cell death was further reduced as the

protein levels resumed to normal levels 72 and 96 hours post doxycycline

induction (Fig. 3-1B). Reduction in STAT3 protein levels also corresponded with

a simultaneous reduction in downstream proteins like Mcl-1 that are under the

transcriptional control of STAT3. We also confirmed the role of STAT3 signaling

in CLL cell viability by using a pharmacological inhibitor of STAT3 called Stattic.

79

Stattic is a non-peptidic small molecule selective inhibitor of STAT3 [198]. It

inhibits STAT3 signaling by selectively inhibits activation, dimerization, and

nuclear translocation of STAT3 [198]. We observed that treatment with Stattic for

24 hours caused a dose-dependent reduction in cell viability in two different CLL

cell lines (Fig. 3-1C (i) and (ii)). JVM-3 is a CLL cell line with wild-type p53, while

Mec-2 cells have mutated p53. Furthermore, we tested three different CLL

patient samples and observed a similar reduction in cell viability after treatment

with Stattic, whereas PMBCs from normal blood donors were resistant to

treatment with Stattic (Fig. 3-1C (iii)). Taken together, these results demonstrate

that STAT3 is essential for CLL cell survival. STAT3 signaling inhibited by either

molecular or pharmacological interventions reduced cell viability and induced

death in CLL cells.

Figure 3-1A. STAT3 is overexpressed in CLL cell lines and patient cells. JVM-3 cells, Mec-2

cells, PBMCs from normal blood donors (n=2) and PBMCs from CLL patients (n=4) were lysed

and protein extracted. Western blot analysis was one to determine the level of total STAT3 in the

samples`

80

Figure 3-1B. Knockdown of STAT3 induces cell death in CLL cells. JVM-3 cells

were transfected with several clones of STAT3 shRNA and flow cytometric analysis was

performed to determine the % dead cells 24-96 hours after induction. Western blot

analysis was done at the same time points to determine the knockdown of STAT3 and

the levels of Mcl-1, a protein regulated by the transcriptional activity of STAT3. Cells

nucleofected with TE buffer containing no plasmid were used as a control. 1µg/mL

doxycycline was used to induce the expression of STAT3 shRNA 24 hours after

nucleofection. This concentration was non-toxic to cells. The results are a mean of two

independent triplicates. Students t-test was used for statistical analysis, *** p < 0.0001,

** p < 0.05.

81

Figure 3-1C. STAT3 inhibition reduces viability of CLL cell lines and patient cells.

(i) JVM-3 cells were treated with increasing concentrations of Stattic for 24 hours and

cell viability was determine after 24 hours treatment. Western blotting analysis for p-

STAT3-Y705, p-STAT3-S727 and total STAT3 was performed to confirm the

effectiveness of 24 hours treatment with Stattic. (ii) Mec-2 cells and (iii) PBMCs from

82

normal donors (PBMC n=2) and CLL patients (CLL n=3) were treated with Stattic for 24

hours and cell viability was determined after treatment.

CNL suppresses the phosphorylation of STAT3 at both tyrosine-705 (p-

STAT3-Y705) and serine-727 (p-STAT3-S727) residues in CLL cells

To identify the mechanism of CNL-induced cell death in CLL, we examined

the effect on STAT3 phosphorylation. JVM-3 cells were treated with CNL or

ghost nanoliposomes (negative control) for 24 hours and Western blotting was

performed to detect changes in protein phosphorylation. While the levels of total

STAT3 protein did not change significantly with CNL treatment, phosphorylation

of STAT3 was significantly down regulated. p-STAT3-S727 and p-STAT3-Y705

were suppressed by 45% and 67% respectively after treatment with 40µM CNL

for 24 hours (Fig. 3-2A (i)). Several studies have reported that only S727 residue

of STAT3 is constitutively phosphorylated in CLL [188, 189]. However, we

observed constitutive phosphorylation at both S727 and Y705 sites in both CLL

cell lines and primary cells. This can be attributed to increased activity of

upstream kinases like cAbl that contribute to STAT3 phosphorylation [199]. We

observed a similar trend on STAT3 phosphorylation suppression after CNL

treatment in Mec-2 cells, although after longer treatment with CNL (Fig. 3-2A (ii)).

Similarly, 7 CLL patient cells were treated with 40µM CNL or ghost

nanoliposomes for 24 hours and evaluated for STAT3 phosphorylation.

Consistent with the above results, we observed suppression in STAT3

phosphorylation at both the residues in 6 out of 7 patient samples (Fig. 3-2B).

Although we observed a slight reduction in total STAT3 levels in some patient

83

cells after CNL treatment, the overall change in total STAT3 did not reach

statistical significance in most patient cells. We also tested the effect on CNL

treatment on HEK293 cells. CNL treatment did not affect the viability of HEK293

cells (Fig. 3-2C (i)), nor did it affect the phosphorylation of STAT3 (Fig 3-2C(ii)),

thereby demonstrating that this phenomenon is specific to cancer cells.

Figure 3-2A. CNL suppresses the phosphorylation of STAT3 in CLL cell lines. (i)

JVM3 cells were treated with 20µM and 40µM of ghost nanoliposomes and CNL for 24

84

hours. Western blotting analysis was performed to determine protein levels before and after

treatment. The graphs represent the quantification of Western blotting. Statistical analysis

was performed using Student’s t-test, * p < 0.05, ** p < 0.01(ii) Mec-2 cells were treated

with 40µM nanoliposomes for 48hours and 72 hours and protein levels were determined

using Western blotting analysis.

Figure 3-2B. CNL suppresses phosphorylation of STAT3 in CLL patient cells. CLL

patient cells (n=7) were treated with 40µM of ghost nanoliposomes and CNL for 24 hours.

Western blotting analysis was performed to determine protein levels before and after

treatment. The graphs represent the quantification of Western blotting. Statistical analysis

was performed using Student’s t-test, * p < 0.05.

85

Figure 3-2C. CNL does affect cellular viability & STAT3 phosphorylation in HEK293

cells. (i) Cell viability of HEK293 cells was determined after 24 hour treatment with ghost

nanoliposomes and CNL. (ii) Western blotting analysis was performed to determine levels

of STAT3 phosphorylation in HEK293 cells after 24 hours treatment with ghost

nanoliposomes and CNL.

Suppression of STAT3 phosphorylation is specific to STAT3 and C6-

ceramide

We next evaluated the specificity of CNL-induced suppression of STAT3

phosphorylation. In addition to STAT3, STAT1 is also constitutively

phosphorylated on S727 in CLL patient cells [188]. Although strong evidence

supporting the role of STAT1 in CLL is lacking, some investigations have

reported that fludarabine and JAK kinase inhibitors induce apoptosis in CLL cells

and inhibit STAT1 signaling, thereby suppressing the ability of the cells to

respond to growth signals, interferons and cytokines [200, 201]. STAT1

activation is also critical for differentiation of CLL cells in response to Byrostatin 1

[202]. These reports raise a possibility that STAT1 might play a role in

pathogenesis of CLL by impacting the anti-apoptotic signals in CLL. We

86

observed that treatment with CNL does not significantly affect the

phosphorylation of STAT1 (Fig. 3-3A). The other STATs were either not

constitutively active (STAT2 and STAT5) or were not expressed in JVM-3 cells

(STAT4). We next examined if suppression of STAT3 phosphorylation was

specific to C6-ceramide sphingolipid. We tested three other sphingolipids:

dihydro-C6-ceramide, an inactive analog of C6-ceramide; sphingosine and

sphingosine-1-phosphate. None of the other sphingolipids tested decreased

STAT3 phosphorylation (Fig. 3-3B), thereby demonstrating the specificity of C6-

ceramide. All together, these results prove the specificity of CNL-induced

suppression of STAT3 phosphorylation in CLL.

Figure 3-3A. CNL-induced suppression of phosphorylation is specific to STAT3. JVM-3

cells were treated with 40µM ghost nanoliposomes and CNL for 24 hours and Western blotting

analysis was done. A positive control of STAT2 phosphorylation was also used.

87

Figure 3-3B. Suppression of STAT3 phosphorylation is specifically an effect of CNL and

not other sphingolipids. JVM-3 cells were treated with dihrdro-C6-ceramide nanoliposomes or

BSA:sphingosine complex or BSA:S1P complex for 24 hours. Western blotting analysis was

performed.

CNL induces necrotic cell death in CLL cells

We have previously demonstrated that CNL selectively induces caspase

3/7-independent cell death in CLL cells [39]. Using phase contrast microscopy

we had shown that CLL cells treated with CNL resembled a necrotic morphology

[39]. Here, we confirm these findings using flow cytometric analysis for Annexin V

and a viability dye, 7-AAD. Several reports suggest that necrotic cell death is

quantified in the Annexin V-7AAD double positive quadrant [203, 204]. Necrotic

cell death was induced in both cell lines in a dose-dependent and time-

dependent manner after treatment with CNL (Fig. 3-4A (i) and (ii)). Under the

same conditions, ghost nanoliposomes did not have effect on cell death. Recent

evidence suggests that CLL patients with p53 pathway dysfuntion have poor

prognosis due to reduced response to conventional chemotherapies [133, 205].

We observed that cell death in p53mutated Mec-2 cells was induced after a longer

88

treatment with CNL as compared to p53wild-type JVM-3 cells (Fig. 3-4A). This

preliminary evidence demonstrating the effectiveness of CNL treatment to induce

cell death in Mec-2 cells presents a potential treatment strategy for p53mutated B

malignancies that are in urgent need of better therapeutic approaches. We also

evaluated the effect of CNL on CLL patient cells obtained from 7 different

patients. Treatment with 40µM CNL for 24 hours induced cell death in 5 out of

the 7 patients tested (Fig. 3-4B). Taken together, these results indicate that CNL

increases cell death in both p53mutated and p53wild-type CLL cells. To determine if

suppression of phosphorylation preceded induction of cell death, we examined

phosphorylation levels at early time points after CNL treatment. Significant

necrotic cell death in JVM3 cells was observed 12 hours after treatment with CNL

(Fig 3-4C (i)). However, suppression of STAT3 phosphorylation at both the

residues started about 6 hours after CNL treatment and thus preceded induction

of cell death (Fig. 3-4C (ii)). This suggests that reduction in STAT3

phosphorylation might mediate CNL-induced cell death. Consistent with this, we

also observed suppression of STAT3 phosphorylation in 3 CLL patient cells after

12 hours of treatment with CNL (Fig. 3-4D). Overall, these results indicate that

CNL suppresses STAT3 phosphorylation and this might mediate cell death in

CLL cells.

89

Figure 3-4A. CNL induces necrotic cell death in CLL cell lines (p53 wt and p53 mutated). (i)

JVM-3 cells that have wild type p53 and (ii) Mec-2 cells that have mutated p53 were treated with

20µM and 40 µM ghost nanoliposomes and CNL for indicated time periods. Flow cytometric

analysis for Annexin V and 7AAD stating was performed to determine effect on cell death. Two-

way ANOVA was used to perform statistical analysis * p < 0.01.

Figure 3-4B. CNL induces necrotic cell death in CLL patient cells. CLL patient cells (n=7)

were treated with 40 µM ghost nanoliposomes and CNL for 24 hours. Flow cytometric analysis for

Annexin V and 7AAD stating was performed to determine effect on cell death. Student’s t test

was used to perform statistical analysis ** p < 0.01.

90

Figure 3-4C. CNL-induced suppression of p-STAT3 precedes induction of cell death (i)

JVM-3 cells were treated with ghost nanoliposomes and CNL for indicated time periods and flow

cytometric analysis was performed to determine % cell death. (ii) JVM-3 cells were treated with

40µM ghost nanoliposomes and CNL for indicated time periods and Western blotting was

performed for to determine levels of p-STAT3. Graphical representation of protein levels is also

shown. Statistical analysis was done using Student’s t-test * p < 0.05.

91

Figure 3-4D. CNL-induces early time point suppression of p-STAT3 in CLL patient cells.

CLL patient cells (n=3) were treated for 12 hours with 40µM ghost nanoliposomes or CNL and

Western blotting was done.

CNL suppresses Bruton’s tyrosine kinase (BTK), mitogen-activated protein

kinase kinase (MEK) and protein kinase C (PKC) activity to suppress STAT3

phosphorylation

Suppression in STAT3 phosphorylation can be a result of CNL-induced

activation of downstream phosphatases and/or CNL-induced suppression in

upstream kinases. We studied the effect of CNL on downstream phosphatases

and upstream kinases to determine which enzymes predominantly mediate CNL-

induced suppression in STAT3 phosphorylation. Okadaic acid (OA) was used as

an inhibitor of serine/threonine phosphatases PP1 and PP2A. As shown in Fig.

3-5A (i), pretreatment with OA did not rescue CNL-induced-suppression of p-

STAT3-S727, indicating that suppression of phosphorylation is not a result of

CNL-induced activation of serine/threonine phosphatases. Pretreatment with OA

rescued p-Akt-S473 after CNL treatment, suggesting that the inhibitor was

functional in inhibiting PP2A and PP1. Similarly, we observed that CNL treatment

92

had no effect on the activity of tyrosine phosphatases. Pervanadate (PV) was

used as a functional inhibitor of tyrosine phosphatases. As demonstrated in Fig.

3-5A (ii), pretreatment with PV did not rescue CNL-induced-suppression of p-

STAT3-Y705, indicating that this event is independent of the action of tyrosine

phosphatases. The basal levels of p-STAT3-Y705 increased after pretreatment

with PV confirming that PV was effective in inhibiting tyrosine phosphatases.

Figure 3-5A. CNL does not activate phosphatases. (i) JVM-3 cells were pretreated with 5nM

okadaic acid (OA) for 2 hours, followed by 12 hours of treatment with ghost nanoliposomes and

CNL. (ii) JVM-3 cells were pretreated with 50µM pervanadate (PV) for 2 hours, followed by 24

hours treatment with 40µM ghost nanoliposomes and CNL. Both the inhibitors were non-toxic to

cells at the specific concentration. Western blotting was performed.

After investigating the effect of CNL on downstream phosphatases, we

studied the effect of CNL on upstream kinases. We used two strategies to

determine the effect of CNL on upstream kinases. Firstly, we performed western

blot analysis to study the effect of CNL on the activating phosphorylation of the

kinase or to look at the phosphorylation pattern of the immediate downstream

target of the kinase. These results act as a surrogate for determining the effect

93

on the kinase activity of the enzyme. Additionally, we also treated cells with an

inhibitor of the kinase to determine the direct effect of kinase inhibition on STAT3

phosphorylation. One of our criteria for screening was to obtain inhibition of

kinase activation at early time points after treatment with CNL, which would

thereby indicate that kinase activity suppression preceded suppression in STAT3

phosphorylation. Taken together, these two strategies answer the question of the

effect of CNL on upstream kinases.

We observed that CNL suppresses the activity of BTK, a tyrosine kinase

critical in mediating BCR signaling in CLL cells. As shown in Fig. 3-5B, we

observed a significant reduction in phosphorylation of BTK at Y223 in JVM-3

cells as early as 4-6 hours after treatment with CNL but not ghost liposomes. The

levels of total BTK remained unchanged after treatment. BTK phosphorylation at

Y223 is an activating phosphorylation and is necessary of full activation of BTK

[206, 207]. CNL-induced inhibition of BTK is an exciting observation since BTK is

a promising target in CLL and a BTK inhibitor, ibrutinib, is currently used in the

clinic for CLL therapy [208]. Furthermore, we also observed that treatment with

ibrutinib significantly reduced p-STAT3-Y705, but not p-STAT3-S727 (Fig. 3-5C).

Taken together, these results imply that CNL-induced BTK inhibition mediates

suppression of p-STAT3-Y705.

94

Figure 3-5B. CNL suppresses the activity

of BTK. JVM-3 cells were treated with 40µM

ghost liposomes and CNL for indicated time

periods and Western blotting was performed.

Figure 3-5C. BTK inhibitors suppress phosphorylation of STAT3. JVM-3 cells were treated

with BTK inhibitor, ibrutinib for 6 hours and Western blotting was performed to determine protein

levels. Graphical representation of the Western blot is also shown. Student’s t-test was used to

perform statistical analysis, * p < 0.05.

We also observed that CNL suppresses MEK activity, a serine/threonine

kinase in CLL cells. As shown in Fig. 3-5D, only CNL treatment significantly

suppressed phosphorylation of Erk, a direct downstream target of MEK at early

time points after addition of the liposomes. Furthermore, treatment with U0126, a

MEK inhibitor reduced p-STAT3-S727 and p-STAT3-Y705 in both JVM-3 (Fig. 3-

95

5E (i)) cells and CLL patient cells (Fig. 3-5E (ii)). We also observed that CNL

suppresses PKC activity. CNL treatment, but not ghost liposomes, significantly

suppressed phosphorylation of MARCKS, a direct downstream target of PKC at

early time points, while total MARCKS levels remained unchanged (Fig. 3-5F).

Additionally, treatment with BisI, a PKC inhibitor also suppressed p-STAT3-S727

and p-STAT3-Y705 levels in both JVM-3 cells and CLL patient cells (Fig. 3-5G (i)

and (ii)).It was confounding to obtain inhibition of p-STAT3-Y705 after treatment

with U0126 and BisI, both of which are inhibitors of serine/threonine kinases.

However, some results in the literature have demonstrated that MEK inhibitors

like U0126 and PKC inhibitors like BisI inhibit p-STAT1-Y701 [202]. Since STAT1

and STAT3 pathways share similar upstream signaling, we speculate that MEK

and PKC might also play a role in phosphorylating STAT3-Y705. We conclude

that CNL-induced suppression in phosphorylation of STAT3 is not mediated by

downstream phosphatases, but instead by inhibition of upstream kinases that

include BTK, MEK and PKC.

Figure 3-5D. CNL suppresses the activity of MEK1/2 kinase. JVM-3 and Mec-2 cells were treated with 40µM ghost liposomes and CNL for indicated time periods and Western blotting was performed.

96

Figure 3-5E. MEK1/2 inhibitors suppress phosphorylation of STAT3. (i) JVM-3 cells were

treated for 6 hours and 12 hours with 10µM U0126 and Western blotting was performed to

evaluate protein levels. Blots were probed for p-Erk to confirm that U0126 suppressed MEK

activity. Graphical representation of the blots is also shown. (ii) CLL cells (n=3) were treated for

12 hours with 10µM U0126 and Western blotting was performed to evaluate protein levels.

Graphical representation of the blots is also shown. Student’s t-test was used for statistical

analysis, * p < 0.05.

97

Figure 3-5F. CNL suppresses the

activity of PKC. JVM-3 cells were

treated with 40µM ghost liposomes

and CNL for indicated time periods

and Western blotting was

performed.

Figure 3-5G.

PKC inhibitor

suppress

phosphorylati

on of STAT3

(i) JVM-3 cells

were treated

for 6 hours and

12 hours with

5µM Bis-I and

Western

blotting was

performed to

evaluate

protein levels.

Graphical

representation

of the blots is

also shown.

(ii) CLL cells

(n=3) were

treated for 12

hours with 5µM

Bis-I and Western blotting was performed to evaluate protein levels. Graphical representation of

the blots is also shown. Student’s t-test was used for statistical analysis, * p < 0.05.

98

CNL suppresses the transcriptional activity of STAT3

Having established that CNL suppresses STAT3 phosphorylation, we next

sought to determine if CNL suppressed the transcriptional activity of STAT3.

CNL treatment caused a significant downregulation of STAT3-regulated proteins

like Mcl-1, survivin, XIAP, cyclin D1 and p21 in both JVM-3 and Mec-2 cells (Fig

3-6A). CNL-induced downregulation of critical anti-apoptotic proteins like Mcl-1,

survivin and XIAP was encouraging since ample evidence in the literature has

established that these anti-apoptotic proteins are critical mediators of CLL cell

survival. Mcl-1 levels correlate with poor disease prognosis and chemoresistance

[209, 210]. Additionally, Mcl-1 inhibitors induce apoptosis in CLL cells, thus

making it a relevant therapeutic target [211]. Similarly, survivin levels have been

correlated to poor prognosis in CLL and targeting survivin with a small molecule

inhibitor effectively induces apoptosis in the proliferative subset of CLL [212,

213]. Abnormally high expression of XIAP has also been observed in CLL, thus

conferring resistance from apoptosis in CLL cells [214]. Additionally, XIAP

inhibitors synergize with death receptor ligand tumor necrosis factor–related

apoptosis-inducing ligand (TRAIL) and anti-CD40 therapies such as CD154 gene

therapy to induce apoptosis in CLL [214, 215]. As shown in Fig 3-4C (ii), CNL-

induced suppression of STAT3 phosphorylation started about 6 hours after

treatment. This event preceded reduction in Mcl-1 levels which started

approximately 6-8 hours after CNL treatment (Fig 3-6B). We confirmed these

results using a luciferase reporter assay, wherein we observed a significant

dose-dependent reduction in luciferase units after 12 hours treatment with CNL

99

(Fig. 3-6C). This suggests that CNL suppresses the transcriptional activity of

STAT3.

Figure 3-6A. CNL reduces the

levels of STAT3-regulated

genes. JVM-3 cells and Mec-2

cells were treated with 20µM or

40µM ghost liposomes or CNL

and Western blotting was

performed. JVM3 cells were

treated for 24 hours and Mec-2

cells were treated with 48 hours.

Figure 3-6B. Reduction of STAT3 phosphorylation precedes reduction of Mcl-1 levels

following CNL treatment. JVM-3 cells were treated with 40µM ghost liposomes or CNL for

indicated time periods and Western blotting was performed for determining Mcl-1 levels.

100

Figure 3-6C. CNL inhibits expression of luciferase in a STAT3 luciferase reporter assay.

JVM-3 cells were transfected with different luciferase constructs. 12 hours after transfection, cells

were treated with ghost nanoliposomes or CNL for 12 hours and luciferase assay was performed

according to the manufacturer’s protocol.

Overexpression of STAT3-C rescues CNL-induced cell death in CLL cells

To confirm the role of STAT3 in CNL-induced cell death, we

overexpressed STAT3-C, an oncogenic form of STAT3 that mimics STAT3

dimers and thus acts as a constitutively active STAT3 [216]. Overexpression was

obtained using lentiviral transduction and cells expressing STAT3-C. The STAT3-

C overexpression vector has an EGFP sequence as a selectable marker and the

transduced cells were sorted for EGFP and grown as a pure population (JVM3-

STAT3C cells). STAT3-C expression was confirmed by expression of the FLAG-

tag in JVM3-STAT3C cells (Fig. 3-7A inset). Moreover, as shown in Fig 4-7A

inset, JVM3-STAT3C cells have a higher expression of Mcl-1, thus indicating

higher STAT3 transcriptional activity. An overexpression vector expressing RFP

101

was used as a control for the study (JVM3-RFP cells). Wild type JVM-3 cells

were also used as a negative control (WT-JVM3). As shown in Fig. 3-7A, WT-

JVM3 cells undergo necrotic cell death on treatment with CNL for 24 hours.

Similarly, JVM3-RFP cells undergo cell death after CNL treatment. However,

cells expressing STAT3-C were significantly more resistant to treatment with

CNL as compared to WT-JVM3 and JVM3-RFP cells. Since STAT3-C

overexpressing cells were more resistant to CNL-induced cell death, we

concluded that STAT3 mediates CNL-induced cell death in CLL cells. We also

determined STAT3 phosphorylation levels in JVM-3 xenograft tumors obtained from

a subcutaneous CLL mouse model in Balb/c Nu/nu mice injected with ghost

nanoliposomes (n=1) or CNL (n=2) [39]. Consistent with our in vitro data, we observed a

significant reduction of STAT3 phosphorylation in the xenograft tumors injected with CNL

(Fig. 3-7B). Thus CNL reduces STAT3 phosphorylation in vitro and in vivo.

102

Figure 3-7: A) STAT3-C expressing cells are resistant to CNL-induced cell death. Lentiviral

transduction was performed to express STAT3-C in JVM-3 cells. 72 hours after the last

transduction, FACS was performed to obtain a pure population of cells expressing STAT3-C and

JVM-3

A.

B.JVM-3 xenograft

tumors treated

with:

103

the treatments were done. An overexpression construct expressing RFP was used as a negative

control. 72 hours after the last transduction cycle, cells were treated with ghost liposomes and

CNL for 24 hours. Expression of STAT3-C was confirmed by Western blotting and probing for

Flag-tag. Flow cytometric analysis for Annexin V and 7AAD was performed to quantitate %

necrotic cells. Student’s t-test was used for statistical analysis, * p < 0.05. B) In vivo

confirmation of CNL-induced suppression in STAT3 phosphorylation. JVM-3 xenograft

tumors were obtained from a subcutaneous CLL mouse model in Balb/c Nu/nu mice that were

injected with ghost nanoliposomes (n=1) or CNL (n=2) (from Ryland LK, PLoS One, 8(12)).

Western blotting was performed to determine the levels of STAT3 phosphorylation.

Discussion

In the present study, we have identified STAT3 as a molecular mediator of

CNL-induced cell death in CLL. Using p53wild-type JVM-3 cells, p53mutated Mec-2

cells, cells from CLL patients and JVM-3 xenograft tumors from a murine model

of CLL, this study presents in vitro and in vivo confirmation for CNL-induced

suppression of STAT3 phosphorylation. Reduction in STAT3 phosphorylation is

followed by reduction in levels of critical anti-apoptotic proteins like Mcl-1,

survivin and XIAP, and eventually leading to cell death. The only study in the

literature investigating the effect of ceramide on STAT1 and STAT3

demonstrates that exogenous ceramide or accumulation of endogenous

ceramide in human fibroblasts induces STAT activation through tyrosine

phosphorylation of STAT1 and STAT3 via JAK2, MEK/ERK, JNK and p38

kinases [216]. However, the relation between ceramide and STAT3 in context of

cancer cells has never been studied. A large body of work has delineated the

signaling cascades that are targeted by endogenous or exogenous ceramide to

exert its pro-apoptotic effects in cancer cells. Some of these targets include AKT,

104

ERK, survivin, phospholipase D, p38 MAPK and death receptor to name a few

[18]. This study demonstrates the inhibitory effect of CNL on STAT3 signaling by

suppressing STAT3 phosphorylation. Consistent with some reports in the

literature, we have convincingly demonstrated that STAT3 is a potential

therapeutic target in CLL.

Several studies have reported that STAT3 is constitutively phosphorylated

on S727 and not Y705 in CLL [188, 189]. However, we observed constitutive

phosphorylation of both S727 and Y705 residues of STAT3 in both CLL cell lines

and patient cells. Although the studies mentioned above did not use JVM-3 cells

or Mec-2 cells for their investigations, our observations of dual phosphorylation

may be attributed to increased activity of upstream kinases like cAbl that

contribute to STAT3 phosphorylation at the tyrosine residue as speculated by

Allen et al. [199]. We have demonstrated that CNL suppresses STAT3

phosphorylation at both S727 and Y705 as early as 6 hours to 9 hours after

treatment, thereby preceding induction of cell death. Since this phenomenon

occurs relatively early after treatment with CNL, we believe STAT3 mediates

CNL-induced cell death in CLL. We observed that CNL-induced STAT3

dephosphorylation suppresses protein levels of critical anti-apoptotic proteins like

Mcl-1, survivin and XIAP that are essential for CLL cell proliferation and

resistance to apoptosis, thus confirming that CNL suppresses the transcriptional

activity of STAT3. Results from the STAT3 luciferase reporter assay confirm that

CNL suppresses the transcriptional activity of STAT3 even at early time points,

105

thereby indicating that CNL-induced suppression in anti-apoptotic proteins are

effectors of cell death, rather than just a consequence of cell death.

We have identified that CNL inhibits the activity of BTK, MEK and PKC

that leads to suppression of STAT3 phosphorylation. CNL-induced inhibition of

BTK is an exciting observation since BTK is a promising target in CLL. Multiple

reports have demonstrated that BTK is a critical kinase for CLL development

[217]. BTK inhibitor, ibrutinib, is currently used in the clinic for CLL therapy [208].

We have also demonstrated that CNL inhibits the kinase activity of MEK and

PKC. PKC has also been shown to be important for CLL development and it is

believed that PKC inhibitors may be an attractive option for CLL treatment [218].

We performed rescue experiments to definitely confirm that STAT3

mediates CNL-induced cell death. Since STAT3 is expressed at very high levels

in CLL, it was difficult to overexpress wild type STAT3 in JVM-3 cells. Hence, we

expressed STAT3-C, a constitutively dimerized (activated) form of STAT3 which

is produced via substitution of amino acid residues 661 and 663 near the C

terminus with cysteine. Stat3C monomers form disulfide bonds with each other,

leading to constitutive dimerization [219]. Several publications have

demonstrated that STAT3-C is oncogenic in nature [219, 220]. We observed that

cells expressing STAT3-C were more resistant to CNL-induced cell death as

compared to wild type JVM-3 and JVM-3 cells expressing a control vector.

STAT3-C expression thus rendered partial protection to CNL-induced cell death,

thereby confirming that STAT3 mediates CNL-induced cell death in CLL.

106

We have previously demonstrated that ceramide targets the Warburg

effect in cancer cells [39]. Ceramide causes a decrease in the protein levels of

GAPDH, a glycolytic enzyme resulting in decreased glycolysis. Pretreating CLL

cells with pyruvate, the end product of glycolysis rescued CNL-induced cell death

and CNL-induced ATP depletion. Thus, targeting GAPDH is one mechanism by

which CNL inhibits aerobic glycolysis [39]. In the present study we have

demonstrated that CNL suppresses STAT3 phosphorylation, resulting in

subsequent inhibition of STAT3 transcriptional activity. Our preliminary studies

have demonstrated that GAPDH may be partially regulated by the activity of

STAT3, which establishes a potential link between CNL-induced suppression in

STAT3 phosphorylation and CNL-induced inhibition of Warburg effect. Another

possible link may be established on the basis of evidence provided by Demaria

et al [221]. The authors have shown that STAT3 acts as a master regulator of cell

metabolism by activating HIF-1α-dependent aerobic glycolysis [221]. Activated

STAT3 upregulates HIF-1α expression, which in turn induces the transcription of

different genes involved in glycolysis [221]. In addition to modulating the levels of

HIF-1α, STAT3 can cooperate with HIF-1α by binding to its responsive

promoters, ensuring the formation of a transcriptionally active complex. It is

believed that STAT3 and HIF-1α form a feed forward loop that leads to enhanced

aerobic glycolysis [222]. Thus, we speculate that CNL-induced suppression in

STAT3 phosphorylation might suppress HIF-1α expression and subsequently

suppress HIF-1α-dependent aerobic glycolysis. This is another mechanism by

107

which CNL-induced suppression in STAT3 phosphorylation may be linked to

suppression in the Warburg effect.

The enhanced function of constitutive p-STAT3-Y705 as a transcription

factor contributing to oncogenic processes has been extensively studied in

multiple tumor cells. Recently, an extranuclear pro-oncogenic role of constitutive

p-STAT3-S727 was uncovered, giving rise to the concept of protein moonlighting

in STAT3. It was reported that STAT3 associates with complex I and II of the

electron transport chain and is required for optimal mitochondrial respiration

[191]. Additionally, phosphorylation at the S727 of mitochondrial STAT3 was

identified to be essential for its mitochondrial function and for Ras-dependent

oncogenic transformation [223]. Following this revelation, reports have

documented the pivotal role of mitochondria-associated constitutive p-STAT3-

S727 in pathogenesis of breast cancer, pancreatic cancer, murine

myeloproliferative neoplasms and also CLL [190, 224-226]. Having established

the ability of CNL to suppress p-STAT3-S727 in CLL, it will be interesting to look

at the effect on phosphorylation of the mitochondrial STAT3, effect on electron

transport chain and overall mitochondrial respiration in CLL.

This work is significant because it is the first body of evidence

demonstrating that CNL suppresses STAT3 phosphorylation in cancer and that

STAT3 mediates CNL-induced cell death. As far as our knowledge, we are the

first to report that CNL inhibits BTK activity, thereby suppressing BTK signaling.

This work thus opens up a wide avenue of research directed towards examining

ceramide-STAT3 relation in other cancer models and developing novel

108

combination therapeutics that demonstrate synergism between CNL and STAT3

inhibitors, especially in STAT3-dependent cancers like CLL.

109

CHAPTER 4: Effect of nanoliposomal C6-ceramide on mitochondrial bioenergetics and mitochondrial STAT3 in chronic

lymphocytic leukemia

Introduction

The role of STAT3 as a transcription factor is well established. Upon

ligand binding to the receptor, an intracellular cascade is initiated which begins

by activation of Janus kinase (JAK) residing on the receptor’s cytoplasmic

domain. Activation of JAK is followed by recruitment and subsequent JAK-

mediated phosphorylation of STAT proteins on a specific tyrosine residue.

Phosphorylated STATs form homo or heterodimers, translocate to the nucleus,

bind to specific DNA sequences and regulate transcription of downstream genes.

STATs thus act to transmit transcriptional signals from cell surface to the nucleus

[227]. The tyrosine residue essential for STAT activation is conserved amongst

all STATs. Several reports have also identified the existence of other receptor

and non-receptor tyrosine kinases, predominantly from the Src family, that

phosphorylate the tyrosine residue and activate STATs [228]. In addition to the

conserved tyrosine, all STATs excluding STAT2 also have a conserved and

phosphorylable serine residue at the C-terminal end. Serine kinases upstream of

STATs include MAPK, PKC, NLK and IKKε [229]. The role of serine

phosphorylation as an inhibitor or activator of the transcriptional activity of STATs

is ambiguous and is cell type-dependent and stimulus-dependent [229].

Recent evidence has uncovered a novel cellular localization and a novel

noncanonical function of STAT3. Studies have revealed that STAT3 also

110

localizes in the mitochondria and controls cell respiration and metabolism [191].

Specifically, mitochondrial STAT3 has been shown to promote oxidative

phosphorylation and enhance activity of complex I and complex II of the electron

transport chain (ETC) [191]. Phosphorylation of STAT3 at serine 727 (S727) is

essential for this function and the mitochondrial function of STAT3 is independent

of its function as a transcriptional factor in the nucleus [191]. Specifically in

context of cancer, phosphorylation of STAT3 S727 in the mitochondria is

essential for Ras-expressing cells to generate subcutaneous tumors in nude

mice. The authors observed that in addition to increasing the activity of electron

transport complex II, STAT3 expression also enhanced the activity of ATP

synthase, and unexpectedly, increased lactate dehydrogenase and decreased

mitochondrial membrane potential [223]. Thus mitochondrial STAT3

phosphorylation mediates cellular transformation by enhancing aerobic

glycolysis, ETC activity, ATP abundance and inhibiting overproduction of ROS.

The role of mitochondrial p-STAT3-S727 in cellular transformation has been

confirmed in several malignancies like CLL, prostate cancer, breast cancer and

myeloproliferative neoplasms [222]. One report has reported that the MEK-ERK

pathway phosphorylates mitochondrial STAT3 on S727 and is necessary for

RAS-mediated transformation [230]. Gene associated with retinoid interferon

induced cell mortality 19 (GRIM 19), a complex I subunit imports STAT3 into the

mitochondria and recruits it to complex I of ETC. Furthermore, S727 of STAT3 is

critical for STAT3-GRIM19 interaction [231]. In addition to mitochondrial STAT3,

even nuclear STAT3 has been linked to energy metabolism and cell

111

transformation. Demaria et al. reported that in MEFs expressing the STAT3C

form, nuclear STAT3 triggered Hif-1α-mediated induction of aerobic glycolysis

and Hif-1α-independent down-regulation of mitochondrial activity [221]. STAT3

tyrosine phosphorylation is critical for this metabolic switch since inhibition of

STAT3 tyrosine phosphorylation downregulated glycolysis prior to growth arrest

and cell death [221].

Multiple reports have shown that STAT3 is constitutively phosphorylated

at S727 residue in CLL cells. However, the molecular function of p-STAT3-S727

overactivation and its role in the pathogenesis of the disease was unknown.

Capron C et al. have reported that p-STAT3-S727 localizes to the mitochondria

and associates with complex I of the ETC in CLL B cells but not normal B cells.

The authors also identified that the glutathione-dependent antioxidant pathways

protected CLL cells against high cellular ROS levels and maintained p-STAT3-

S727 overactivation, thereby mediating stromal protection and viability of CLL

cells. Taken together, the authors concluded that overactivation of p-STAT3-

S727 provided stress protection and viability to CLL cells and correlated with CLL

cell resistance to apoptosis [232]. More recently, researchers have focused on

studying the metabolism of CLL cells to understand the effect of

microenvironment on the disease and to identify potential therapeutic targets.

CLL cells display oxidative stress and are characterized by have increased

mitochondrial ROS production and adaptation to intrinsic oxidative stress by

antioxidant pathways [233]. Furthermore, it has been reported that unlike other

malignant cells, CLL cells have increased oxidative phosphorylation but not

112

increased glycolysis and have increased mitochondrial biogenesis that might

contribute to CLL oncogenesis [233]. Thus targeting the respiratory chain and

promoting mitochondrial ROS can be potential targets for therapy. However, the

microenvironment, specifically the stromal cells cause a glycolytic switch in CLL

cells in a Notch-c-myc signaling-dependent manner, thereby making glucose

metabolism a therapeutically exploitable target [131].

In this study, we aim at evaluating the effects of CNL on mitochondrial

structure, function and mitochondrial bioenergetics of CLL cells. Ample evidence

in the literature has reported the effect of ceramides on mitochondrial integrity

and function. We aim at extending these studies in CLL cells and evaluating the

role of CNL-induced suppression in STAT3 phosphorylation in perturbing

mitochondrial function and cellular bioenergetics.

Methods and Materials

Reagents

Antibodies specific for STAT3, p-STAT3-Y705, p-STAT3-S727, cytochrome c,

cyclooxygenase IV, tubulin and β-actin were purchased from Cell Signaling

Technology Inc. (Beverly, MA). For Western blotting, 12% precasted Nupage

electrophoresis gels from Invitrogen (Carlsbad, CA), and chemiluminescence

reagent from Amersham Biosciences Inc. (Piscataway, NJ) were obtained.

113

Patient characteristics and preparation of peripheral blood mononuclear

cells

Patient characteristics and preparation of peripheral blood mononuclear cells

have been described in detail in Chapter 3 of the dissertation.

Cell culture

Cell culture methods have been described in Chapter 3 of the dissertation.

Preparation of nanoliposomal ceramide (CNL)

Preparation of nanoliposomal ceramide has been described in Chapter 2 of the

dissertation.

Cell viability assay

Cell viability assays were performed as described in Chapter 2 of the

dissertation.

Assessment of mitochondrial membrane potential (∆Ψm):DiOC6 Staining

To evaluate ∆Ψm, JVM-3 cells and CLL patient primary cells were treated with

ghost or CNL at a concentration of 25 µM for 24 hours. This dose was selected

based on studies in Chapter 2 of the thesis. Fifteen minutes prior to the end of

incubation, cells were labeled with DiOC6 (40 nM in PBS) at 37°C as described

previously [234]. After washing, cells were analyzed by flow cytometry.

114

Determination of ROS Production

Dihydroethidium (DHE) is a lipophilic cell-permeable dye that can undergo

oxidation to ethidium bromide in the presence of superoxide, and, to a lesser

extent, hydrogen peroxide and hydroxyl radicals. Ethidium then binds irreversibly

to the double-stranded DNA, causing amplification of a red fluorescent signal,

indicating ROS production [235]. DHE fluorescence was analyzed by flow

cytometry (excitation 488 nm and emission 585 nm).

Confocal Studies

To verify accumulation of C6-ceramide in JVM-3 cells, we formulated a

nanoliposomal C6-ceramide vesicle with 10 mol% NBD-C6 ceramide as a marker

for C6. Cells were plated at 2.0 × 104/well in 8-well chamber slides and allowed to

grow overnight. Nanoliposomal-NBD-C6 ceramide was treated at 25μM for a 2-

hour treatment period. Cellular nuclei were counterstained with 4′,6-diamidino-2-

phenylindole, and MitoTracker Deep Red 633 (Molecular Probes) was used as a

marker for mitochondria following the manufacturer's instructions. C6-ceramide

delivery and accumulation was evaluated by confocal microscopy at 63×

magnification (Leica Microsystems).

Extracellular flux assay

Bioenergetics of JVM-3 cells was determined using the XF96e Extracellular Flux

Analyzer (XF Cell Mito Stress Test Kit, Seahorse Bioscience, North Billerica,

MA). Cells were seeded in specialized tissue culture plates at a concentration of

600,000 cells per well in XF media with 25mM glucose onto a Cell-Tak (BD

115

Biosciences)-coated XF96 plate. Oligomycin (1 μM), carbonyl cyanide 4-

(trifluoromethoxy) phenyl hydrazone (2 μM), rotenone (1 μM), and antimycin (1

μM) were added sequentially, and measurements of extracellular flux were

recorded. Glycolytic capacity was calculated from extracellular acidification rate

(ECAR) values (an indicator for lactic acid production or glycolysis) after the

inhibition of ATP synthase by the addition of oligomycin, when cells resort to

meeting their energy demands using their maximum glycolytic capacity. ECAR

data were analyzed using Wave software. All experiments were performed in at

least quadruplets.

Mitochondrial Protein isolation

Mitochondrial protein isolation was performed using the Mitochondria Isolation kit

for cultured cells (Thermo Scientific, Waltham, MA). 20 million cells per sample

were used for isolation of the mitochondrial protein. Cells were added in reagent

A and incubated on ice for 2 minutes. 10µL of reagent B was added and cells

were vortexed. Reagent C was then added to the cells and after a series of

centrifugations, the supernatant contained cytosolic proteins, while the pellet

contained intact mitochondria. To extract mitochondrial proteins, the pellet was

washed in Reagent C and resuspended in 2% CHAPS in TBS. Mitochondrial

protein was obtained after a short centrifugation spin. Protein assay (DC protein

assay, Biorad) was performed to determine the protein concentration of cytosolic

and mitochondrial proteins.

116

Statistical analysis

All data are expressed as mean plus or minus SEM. Paired Student t test (2-tail

paired) and 2 way analysis of variance test were used to determine the statistical

significance and P value of 0.05 or less was considered statistically significant.

Results

CNL increases mitochondrial ceramide, intracellular ROS and decreases

mitochondrial membrane potential

Mitochondrial damage and dysfunction are associated with programmed cell

death and it has been demonstrated that ceramide accumulates and acts on the

mitochondria to initiate antiproliferative effects [159, 236]. Therefore, we

hypothesized that CNL treatment would have a similar effect and also result in

an accumulation of ceramide in the mitochondria of CLL cells. We observed that

25 µM nanoliposomal-NBD-C6-ceramide colocalized with a MitoTracker Deep

Red marker for mitochondria, signifying accumulation of exogenous ceramide in

the mitochondria in CLL (Fig. 4-1 (i)). To determine whether mitochondrial

membrane integrity was damaged by CNL, mitochondrial membrane potential

(∆Ψm) was measured using DiOC6 staining. As shown in Fig. 4-1 (ii), treatment

with 25 µM CNL for 24 hours induced loss of membrane potential in JVM3 cells,

and also in leukemic cells from a patient with CLL. These studies indicate that

CNL localizes within the mitochondria and disrupts the mitochondrial membrane

117

potential in CLL, which could potentially contribute to mitochondrial dysfunction

and resultant cell death.

Ceramide has also been suggested to act on mitochondria and generate

reactive oxygen species (ROS) [237]. ROS can activate apoptosis, but recently

have been reported to also induce autophagy and other cell death processes.

Moreover, ceramide treatment has also been linked necrotic cell death via rapid

formation of ROS. Treatment of A20 B-cell lymphoma cells with exogenous cell

permeable, C6-ceramide, induced necrotic cell death in a dose- and time-

dependent fashion. This was associated with formation of ROS, and was

accompanied by dissipation of mitochondrial membrane potential [36]. In

addition, presence of ROS scavengers, like Tiron, blocked ceramide-induced

necrosis. Ceramide accumulation in these cells was also associated with a

significant depletion of intracellular ATP levels which indicated induction of

necrotic death [37]. Therefore, we asked whether C6-ceramide treatment could

induce ROS production in CLL cells. Indeed, Fig. 4-1 (iii) demonstrated that

treatment of 25 µM CNL for 24 hours led to production of ROS, compared to

treatment with the ghost nanoliposome. In addition, pre-treatment with the

antioxidant and ROS scavenger, N-acetyl-L-cysteine (NAC) was sufficient to

attenuate this increased ROS production associated with ceramide (Fig. 4-1(iii)).

To determine if pre-treatment with NAC was adequate to rescue cell death

induction by CNL, a cell viability assay was performed. It was evident that there

was some rescue of cell death induction when formation of ROS were inhibited

118

(Fig. 4-1 (iv)), suggesting that ROS production was contributing to the cell death

induction process to some degree.

Figure 4-1. CNL treatment results in accumulation of ceramide in the mitochondria,

decreased mitochondrial membrane potential, and generation of ROS A). Confocal

microscopic image of 25 µM (24 hour treatment) nanoliposomal-NBD-C6-ceramide delivery to

JVM3 cells showing that NBD-C6-ceramide (green) colocalized with cellular mitochondria (red);

Areas of co-localization (overlay) were characterized by appearance of the orange color. Cellular

nucleas was stained with Hoechst 33342 (blue). Magnification, 63X. B). DiOC6 was used to

determine changes in mitochondrial membrane potential using fluorscence-activated cell sorter

analysis. After treatment with 25 µM ghost nanolipsomes (black) or CNL (grey) for 24 hours,

JVM3 and CLL patient cells were collected and stained with DiOC6. C6-ceramide induced a shift

to the left in DiOC6 fluorescence, indicative of an impairment in the mitochondrial membrane

potential. C). Dihydroethidium (DHE) was used to determine ROS production. After treatment with

25 µM ghost nanoliposomes (top purple tracing) or CNL (red tracing) nanolipsomes for 24 hours,

JVM3 cells were collected and stained with DHE. CNL induced a shift to the right in ethidium

119

fluorescence, indicative of increased ROS production. Pre-treatment with 1mM of the anti-oxidant

N-acetyl-L-cysteine (NAC) was able to attenuate this increase in ROS production. D). JVM3 cells

pre-treated with 1mM of NAC then treated for 24 hours with various doses of ghost or CNL. MTT

assay was then performed and percentage of non-viable cells was determined.

CNL treatment causes release of apoptosis inducing factor (AIF) from

mitochondria into the cytosol

It is well established that in conjunction with proapoptotic proteins like BAX,

ceramide forms channels in the mitochondrial membrane resulting in

mitochondrial outer membrane permeabilization (MOMP) [18]. MOMP causes

release of apoptogenic factors like cytochrome c and AIF from the mitochondrial

intermembrane space into the cytosol. While we have already demonstrated that

treatment with CNL does not cause release of mitochondrial cytochrome c into

the cytoplasm and does not cause apoptotic cell death in JVM-3 cells [39], we

wanted to evaluate the effect of CNL on AIF release. AIF released into the

cytosol moves to the nucleus and promotes nuclear chromatin condensation and

DNA fragmentation. Several reports in the literature have shown that ceramide

causes release of mitochondrial AIF into cytosol thereby inducing cell death [238-

241]. As shown in Fig. 4-2, we observed release of AIF into the cytosol in JVM-3

cells after treatment with CNL but not ghost nanoliposomes. AIF release has also

been linked to necrosis/necroptsis, which is the predominant form of cell death

observed in CLL cells after CNL treatment [39, 242].

120

Figure 4-2. CNL treatment results in cytosolic release of AIF. JVM-3 cells were treated with

40µM CNL or ghost nanoliposomes for 24 hours. Western blot analysis was performed on

cytosolic and mitochondrial fractions for AIF, cytochrome c (cyto c), tubulin and cyclooxygenase

IV (Cox IV). Cyto c and cox IV are used as mitochondrial markers and tubulin is used as a

cytosolic marker. Absence of cox IV and tubulin in the cytosolic and mitochondrial fractions

respectively indicate purity of the fractions. Results from two independent experiments are shown

here.

CNL diminishes the transport of STAT3 inside the mitochondria and

subsequent STAT3 phosphorylation

We have previously observed that CNL diminishes phosphorylation of

STAT3 in both CLL cell lines and patient cells. Given the critical role of

mitochondrial p-STAT3-S727 in CLL, we next evaluated the effect of CNL on

mitochondrial STAT3. We observed that treatment with CNL but not ghost

liposomes for 14 hours decreased total STAT3 levels in the mitochondrial fraction

(Fig. 4-3). This phenomenon was more pronounced 24 hours after treatment with

121

CNL, as we observed a complete depletion of total STAT3 in the mitochondrial

fraction of cells (Fig. 4-3). There was a reduction in the levels of phosphorylated

forms of STAT3, p-STAT3-S727 and p-STAT3-Y705 in the mitochondrial fraction

of cells after treatment with CNL. The fact that we observed a reduction in total

mitochondrial STAT3 levels at an early time point, i.e. 14 hours after treatment

with CNL indicated that this phenomenon was not just a consequence of

extended treatment times or cell death, but could potentially affect mitochondrial

bioenergetics and could play a role in CNL-induced cell death. These results

were interesting since it indicated that the effect of CNL on cellular bioenergetics

could be a potential mechanism of cell death in CLL cells.

Figure 4-3. CNL treatment dimishes levels of total STAT3 and p-STAT3 in the

mitochondria. JVM-3 cells were treated with 20 µM or 40µM CNL or ghost nanoliposomes for 14

hours and 24 hours. Western blot analysis was performed on cytosolic and mitochondrial

fractions for p-STAT3-Y705, p-STAT3-S727, STAT3,cyclooxygenase IV (Cox IV) and tubulin. Cox

IV is used as a mitochondrial marker and tubulin is used as a cytosolic marker. Absence (or

122

highly reduced levels) of cox IV and tubulin in the cytosolic and mitochondrial fractions

respectively indicate purity of the fractions. Above are representative blots from two independent

experiemnts.

Effect of CNL on cellular bioenergetics in CLL

To study the effect of CNL on cellular bioenergetics, we first performed

experiments to evaluate glycolysis in cells before and after treatment with CNL.

Cellular glycolysis can be measured by quantifying the lactic acid production via

the extracellular proton release (extracellular acidification rate [ECAR]). As

shown in Fig. 4-4, JVM-3 cells a significant reduction in glycolytic capacity after

treatment with CNL. Treatment with 20µM CNL for 6 hours did not significantly

affect ECAR. This corresponds with the fact that significant cell death is not

observed 6 hours after treatment with CNL. However, treatment with 20µM or

40µM CNL for 12 hours or 24 hours significantly reduced the glycolytic capacity

of JVM-3 cells as compared to when treated with ghost liposomes. These results

indicate that CNL treatment suppresses the glycolytic capacity of CLL cells.

These results also serve to validate our conclusion from Chapter 1 of the thesis,

i.e. CNL suppress the Warburg effect in CLL cells.

Figure 4-4. CNL

treatment suppresses

the glycolytic flux of

JVM-3 cells. JVM-3 cells

were treated with 20 µM or

40µM CNL or ghost

nanoliposomes for 6

hours, 14 hours and 24

hours. Mito stress kit was

123

used to measure ECAR in cells treated with ghost liposomes or CNL. Each sample was run in

four replicates. Error bars show mean +/- SEM. ** p ≤ 0.005 (Student’s t-test).

Discussion

Cellular bioenergetics is a crucial system for maintaining cell viability and its

perturbation is an important component of several cell death pathways. In

addition to being the powerhouse of a cell, the mitochondrion is also an effector

organelle of some cell death pathways and a convergence point for other death

mechanisms. The mitochondrion is critical in death because it functions like a

“ticking bomb” once a death stimulus is received by the cell. Firstly, it sequesters

apoptogenic factors in the intermembrane space that can be released under the

influence of certain stimuli. Furthermore, through oxidative phosphorylation, it is

also a dominant source of ROS that can act as a major mediator of cell death.

We investigated the effect of CNL on mitochondrial integrity of CLL cells. As

expected, we observed that CNL treatment resulted in preferential accumulation

of ceramide in the mitochondria. Several reports in the literature have

documented ceramide localization within the mitochondria after treatment with

CNL [189]. Consistent with other reports, we also observed a reduction in

mitochondrial membrane potential, and increased MOMP after treatment with

CNL. This observation is supported by extensive reports that have demonstrated

that ceramide molecules form macrodomains and channels in the mitochondrial

membrane, thereby compromising MOMP and mitochondrial membrane potential

[18]. CNL-induced increased MOMP also released the pro-apoptotic protein, AIF,

which causes DNA fragmentation. Thus, in line with several other reports, we

124

observed that CNL perturbs mitochondrial structure in CLL. Our next step was to

investigate the effect on mitochondrial function and bioenergetics.

Cancer cells are characterized by metabolic reprogramming, i.e.

dependence on aerobic glycolysis or Warburg effect for ATP production rather

than the more efficient oxidative phosphorylation [233]. This glycolytic switch has

emerged as an attractive target in novel chemotherapeutic strategies. The two

papers published in 2009 that revealed the vital role of mitochondrial p-STAT3-

S727 in cellular respiration and in Ras-mediated cellular transformation through

increased oxidative phosphorylation have been seminal in this field of cellular

bioenergetics [191, 223]. Researchers are now focusing their efforts towards

gaining deeper insights into the dynamics between the Warburg effect and

oxidative phosphorylation for ATP synthesis in tumor cells.

Mitochondrial STAT3 phosphorylation mediates cellular transformation by

enhancing aerobic glycolysis, ETC activity, ATP abundance and inhibiting

overproduction of ROS [222]. p-STAT3-S727 is particularly important in CLL

because it is constitutively active in all CLL cells irrespective of the

phosphorylation status of Y705 [188]. Similar to other malignancies, p-STAT3-

S727 localizes to the mitochondria and associates with complex I of the ETC in

CLL cells [232]. Overactivation of p-STAT3-S727 provides stress protection and

viability to CLL cells and correlates with CLL cell resistance to apoptosis,

indicating that its potential function as an enhancer of oxidative phosphorylation

and/or glycolysis might be crucial for CLL cell viability and chemoresistance

[232]. In support of this speculation, recent reports indicate that unlike other

125

malignancies, CLL cells have increased oxidative phosphorylation but not

increased aerobic glycolysis [233]. However, stromal cells have been shown to

cause a glycolytic switch in CLL cells [131].

Since we have previously reported that CNL treatment suppresses levels

of p-STAT3-S727 in CLL cells (Chapter 3), we next investigated the effect of CNL

specifically on mitochondrial STAT3. We observed that CNL suppressed

mitochondrial STAT3 phosphorylation, but more importantly we noticed reduced

levels of total STAT3 in the mitochondria of cells treated with CNL. Several

possibilities can explain this observation. Since STAT3 shuttles between the

mitochondria and cytoplasm, it is likely that dephosphorylation of mitochondrial

STAT3 at S727 enhances its export out of the nucleus, thus causing a reduction

in total STAT3 levels in the mitochondria. Future work in this scenario should

include identifying the mitochondrial kinases or phosphatses regulated by

mitochondrial ceramide to cause this suppression in p-STAT3-S727 and the

eventual export of STAT3 out of the mitochondria. Another possibility is CNL-

induced reduction in the mitochondrial import of STAT3. GRIM-19 is known to

shuttle STAT3 into the mitochondria and phosphorylation of S727 residue of

STAT3 is critical for this interaction [231]. Hence, the effect of CNL on GRIM-19

levels and function is another possibility that should be investigated.

Next, we directly investigated the effects of CNL on cellular bioenergetics,

i.e. aerobic glycolysis and oxidative phosphorylation. Lactic acid release in the

extracellular environment (ECAR) is a measure of glycolysis and glycolytic flux.

We observed a time-dependent reduction in the glycolytic flux of CLL cells after

126

treatment with CNL. This result validates our previously published report that

CNL targets the Warburg effect in CLL cells by reducing levels of GAPDH

(Chapter 2) [39]. It is well accepted that the glycolytic switch is an attractive

target for novel chemotherapeutics. In CLL, targeting the Warburg effect can

prove to be promising, as this strategy will also overcome microenvironment-

mediated CLL cell viability and chemoresistance [131]. We are currently

investigating the effects of CNL on oxidative phosphorylation of CLL cells. We

speculate that CNL treatment will also inhibit oxidative phosphorylation in CLL

cells, since we have consistently observed a reduction in mitochondrial p-STAT3-

S727 on treatment with CNL. Our current and future work includes studying the

effect of CNL on cellular bioenergetics in CLL cells transfected with mitochondria-

targeted STAT3 constructs mutated at S727 and/or Y705 residues. These

experiments will provide a clear picture of the role of mitochondrial p-STAT3-

S727 in CNL-induced alterations in cellular bioenergetics and cell death.

Taken together, this work is crucial for uncovering the exact effect of CNL

on cellular bioenergetics in CLL. This research can identify potential therapeutic

targets in the cellular bioenergetics system that can also be inhibited to improve

the efficiency of CNL as a therapy for CLL.

127

CHAPTER 5 – Conclusions, Future directions and Therapeutic Potential

Conclusions

This dissertation has broadly focused on elucidating the molecular

mechanisms of C6-ceramide-induced cell death in chronic lymphocytic leukemia.

C6-ceramide is delivered in a nanoliposomal formulation to overcome the

delivery challenges due to the hydrophobicity of the sphingolipid. It is well

established that CNL is a lucrative delivery formulation for ceramides in several

cancer models and CNL suppresses tumor growth in models of breast cancer,

sarcoma, melanoma, hepatocellular carcinoma and LGL leukemia [18]. This work

aims at testing the use of CNL as a therapy for CLL and also to elucidate the key

molecular signaling pathways mediating CNL-induced cell death in CLL.

Following are the major conclusions presented in this dissertation:

Use of nanoliposomal C6-ceramide as a therapy for CLL

Sphingolipid-based therapeutics is a good strategy for treating CLL. This

conclusion is supported by reports in the literature that demonstrate a

dysregulation of sphingolipid metabolism in CLL. A recent report has uncovered

a novel link between BCR signaling and sphingolipid metabolism. It has been

reported that BCR controls chemoresistance of primary CLL cells by controlling

glucosylation of ceramides via upregulation of UGCG, thereby reducing levels of

proapoptotic ceramide species [139]. Use of BCR signaling inhibitors inhibit

UGCG resulting in an increase intracellular ceramide levels and subsequent

128

chemosensitization of CLL cells [139]. Another report has demonstrated that

sphingolipid metabolism is a potential novel mechanism of CLL [140].

We have demonstrated the effectiveness of CNL in CLL cells both in vitro

and in vivo. We observed that CNL induces caspase-independent, non-

apoptotic, necrotic cell death in CLL cell lines and CLL patient cells. Given the

fact that CNL was also effective in inducing cell death in Mec-2 cells which have

mutated p53, we hypothesize that this therapy, stand-alone or in combination,

would be effective against high risk disease groups like CLL with p53 mutation

cytogenetics.

We obtained promising in vivo evidence that CNL demonstrates in vivo

efficacy via tumor growth inhibition in a murine xenograft model of CLL. We are

aware of a recently described mouse model for CLL generated by co-injection of

primary human CLL cells and T cells into NOD/ SCID mice, but have chosen to

focus our studies on the mouse xenograft model because we can obtain tissue

for ex vivo Western analysis of GAPDH protein levels [182]. In our in vivo studies

we observed that treatment with nanoliposomal C6-ceramide effectively

decreased tumor burden without systemic side effects. Although necrosis is not

the preferred cell death mechanism, as it elicits immune and inflammatory

response that can be detrimental, we did not observe any systemic side-effects

in our animal study. In addition, it was expected that this nanoliposomal

formulation would be relatively non-toxic to the animals, as it was previously

selected by the Nanotechnology Characterization Laboratory (National Cancer

Institute) for extensive toxicology and stability testing. (Detailed information on

129

the toxicology studies of the “Ceramide Liposomes” can be found at

(http://ncl.cancer.gov/working_technical_reports.asp).

CNL targets the Warburg effect in CLL

Cancer cells are characterized by a glycolytic switch, i.e. preference for

aerobic glycolysis or Warburg effect for energy production over the more efficient

oxidative phosphorylation. Several studies have concluded that the Warburg

effect is a therapeutic target for treating cancer and new drugs are being

developed to target different steps of this phenomenon [243]. Specifically in CLL,

a recent paper reported that stromal cells promote a glycolytic switch in CLL cells

and targeting glucose metabolism can be exploited to breach stromal cell-

mediated drug resistance [131]. We have demonstrated that CNL decreases

protein and mRNA expression of GAPDH, an enzyme that oxidizes

glyceraldehyde-3-phosphate to 1,3-biphosphoglycerate. GAPDH is a potent

inhibitor of caspase-independent cell death by enhanced glycolysis and

enhanced autophagy [167]. Several reports have revealed that overexpression of

GAPDH may assist in conferring chemoresistance to cancer cells [166, 175]. We

observed that CNL targets GAPDH-dependent glycolysis. This was further

confirmed by observations like CNL-induced reduction in ATP, lactic acid and

rescue of ATP depletion and cell death by pretreatment with pyruvate which is a

final product of glycolysis. We did not see any changes in glucose uptake or

changes in GLUT1. This indicated that CNL directly targets the glycolytic

pathway. However, the mechanism of CNL-induced GAPDH inhibition is still

unknown. We provide evidence that the glycolytic pathway is an attractive target

130

in CLL that should be explored and our data emphasizes the potential

importance of GAPDH as a target of CNL.

STAT3 mediates CNL-induced cell death in CLL

STAT3 is a well-established therapeutic target in several cancers. Several

reports have demonstrated that STAT3 is constitutively phosphorylated on S727

in CLL cells independent of the phosphorylation status of Y705 residue [188,

189]. p-STAT3-S727 in CLL cells binds to DNA and activates transcription in CLL

cells irrespective of phosphorylation at Y705 [189]. Researchers have long

speculated that STAT3 might be a therapeutic target in CLL and some STAT3

inhibitors have been explored for therapy in CLL [189, 196, 197]. Our data

demonstrates that CNL suppresses STAT3 phosphorylation at both S727 and

Y705 residues. CNL-induced suppression of STAT3 phosphorylation results in

suppressing the transcriptional activity of CLL, subsequently downregulating anti-

apoptotic proteins like Mcl-1, survivin and XIAP. CNL-induced reduction in these

anti-apoptotic proteins is very promising since these proteins are critical for CLL

pathogenesis [209, 213]. We have also identified the upstream kinases that are

modulated by CNL to cause suppression in STAT3 phosphorylation. We have

demonstrated that CNL inhibits the activity of BTK, MEK and PKC kinases to

suppress STAT3 phosphorylation. CNL-induced suppression of MEK/ERK

pathway is consistent with published reports [35]. The role of CNL on PKC

activity is debatable since some reports provide evidence that C6-ceramide

activates PKC in vascular smooth cells, while some demonstrate that C6-

cermaide suppresses tumor metastasis by eliciting PKCζ tumor-suppressive

131

activities [146, 244]. However, our observation that CNL suppresses PKC activity

in CLL supports reports demonstrating that PKC is critical for CLL development

and a potential druggable target [245]. The inhibitory effect of CNL on the kinase

activity of BTK is novel and promising since BTK is a critical kinase for CLL

development. Ibrutinib, a BTK inhibitor, is currently being prescribed for the

treatment of CLL [208]. Our data imply that CNL might be partially inhibiting the

crucial BCR signaling in CLL cells. Taken together, we report novel data that

STAT3 phosphorylation and BTK are targets of CNL.

Effect of CNL on mitochondrial STAT3

The role of ceramide in regulating mitochondrial integrity during the cell

death processes has been extensively studied [27]. Furthermore, with the

discovery of the novel function of STAT3 and specifically p-STAT3-S727 in

regulating mitochondrial oxidative phosphorylation, the next obvious step is study

the effects on mitochondrial function after CNL treatment [191]. We have shown

that CNL treatment reduces mitochondrial membrane potential, induces MOMP

resulting in subsequent AIF release and increases ROS production. Ceramide

has been previously shown to cause such changes in the mitochondria in

cancerous cells [35]. We have studied the effect of CNL treatment on

mitochondrial STAT3 and we observe that treatment with CNL reduces the total

STAT3 levels within the mitochondria which results in lower levels of

mitochondrial phosphorylated STAT3. Since STAT3 shuttles between the

mitochondria and cytoplasm, it is likely that dephosphorylation of mitochondrial

STAT3 at S727 enhances its export out of the nucleus, thus causing a reduction

132

in total STAT3 levels in the mitochondria. Another possibility is CNL-induced

reduction in the mitochondrial import of STAT3, GRIM-19. We have also

examined the effect of CNL on cellular bioenergetics. We have observed a CNL-

induced time-dependent reduction in the glycolytic capacity of CLL cells, which

further validates our conclusion that CNL targets the Warburg effect in CLL.

Effect of CNL on cellular bioenergetics in CLL

In this thesis project, we have discovered three different targets of CNL in

CLL that include the Warburg effect, STAT3 transcriptional activity and STAT3

mitochondrial function. All three components play a significant role in maintaining

cellular bioenergetics. Thus it is intuitive to explore if there exists a link between

these three targets.

Recent work has uncovered a link between glycolytic metabolism and

sphingolipid metabolism. The authors show that glucose availability and

glycolytic metabolism dictate glycosphingolipid levels [246]. Treatment of

leukemic cells that have elevated glycosphingolipids with inhibitors of glycolysis

or the pentose phosphate pathway significantly decreased GlcCer levels and

increased intracellular ceramide levels leading to sensitization to cytotoxic drugs

[246]. However, there is very little evidence suggesting if the same relation exists

in reverse, i.e. if sphingolipid metabolism affects glycolytic metabolism. One

report demonstrated that ceramide starves cells to death by inducing intracellular

nutrient limitation via downregulation of nutrient transporter proteins [148].

However, the authors do not present data about the effect on nutrient transporter

protein downregulation on glycolysis. In this work we have established a novel

133

relation between ceramide and glycolysis. We have demonstrated that ceramide

targets the Warburg effect in cancer cells. Ceramide decreases the protein levels

of GAPDH, a glycolytic enzyme resulting in decreased glycolysis. We have

shown that pretreating CLL cells with pyruvate, the end product of glycolysis

rescues CNL-induced cell death and CNL-induced ATP depletion. Thus,

targeting GAPDH is one mechanism by which CNL inhibits aerobic glycolysis.

We have demonstrated that CNL suppresses STAT3 phosphorylation. It has

been reported that CLL cells are characterized by STAT3 which is constitutively

phosphorylated at S727 [189]. p-STAT3-S727 binds to DNA and exhibits

transcriptional activity in CLL cells independent of Y705 phosphorylation [189]. In

our model, we show that CNL-induced suppression in STAT3 phosphorylation

inhibits STAT3 transcriptional activity. Our initial studies demonstrated that

GAPDH may be partially regulated by the activity of STAT3, which establishes a

link between CNL-induced suppression in STAT3 phosphorylation and CNL-

induced inhibition of Warburg effect. Another possible link may be established on

the basis of evidence provided by Demaria et al. The authors have shown that

STAT3 acts as a master regulator of cell metabolism by activating HIF-1α-

dependent aerobic glycolysis [221]. Activated STAT3 upregulates HIF-1α

expression, which in turn induces the transcription of different genes involved in

glycolysis [221]. In addition to modulating the levels of HIF-1α, STAT3 can

cooperate with HIF-1α by binding to its responsive promoters, ensuring the

formation of a transcriptionally active complex. It is believed that STAT3 and HIF-

1α constitute a feed-forward loop that enhances aerobic glycolysis [222]. Thus,

134

we speculate that CNL-induced suppression in STAT3 phosphorylation might

suppress HIF-1α expression and subsequently suppress HIF-1α-dependent

aerobic glycolysis. This is another mechanism by which CNL-induced

suppression in STAT3 phosphorylation may be linked to suppression in the

Warburg effect.

Ceramide reduces mitochondrial membrane potential, increases MOMP

and subsequently increases ROS production. Furthermore, short-chain

ceramides have also been shown to inhibit the mitochondrial respiratory function,

ETC and oxidative phosphorylation [247]. Although the exact mechanism of

ceramide-induced inhibition of ETC is not clear, it is speculated that ceramide

macrodomains in the mitochondrial membrane may alter the hydrophobicity of

the membrane thus altering protein-lipid interaction. These biophysical alterations

affect the structure of the supercomplexes of the ETC, thus inhibiting function.

Another possibility is that ceramide may modify the activity of ETC by acting as

an allosteric effector by specifically binding to individual ETC complexes [247].

We are currently conducting experiments to provide direct evidence that CNL

reduces mitochondrial respiration but we speculate a CNL-induced reduction in

mitochondrial respiration. Recent reports have confirmed that mitochondrial p-

STAT3-S727 is required for oxidative phosphorylation and for Ras-mediated

cellular transformation. Inhibition or deletion of p-STAT3-S727 suppresses the

activity of complex I, II and/or V of the ETC [191, 223]. Similar to other

malignancies, p-STAT3-S727 localizes to the mitochondria and associates with

complex I of the ETC in CLL cells [232]. Overactivation of p-STAT3-S727

135

provides stress protection and viability to CLL cells and correlates with CLL cell

resistance to apoptosis, indicating that its potential function as an enhancer of

oxidative phosphorylation and/or glycolysis might be crucial for CLL cell viability

and chemoresistance [232]. We have demonstrated that CNL decreases

mitochondrial STAT3 protein and a subsequent decrease in p-STAT3-S727. We

hypothesize that CNL-induced reduction in p-STAT3-S727 is a major player in

CNL-mediated inhibition of mitochondrial respiration.

Taken together, we conclude that CNL treatment has a pronounced effect

on cellular bioenergetics in CLL cells leading to cell death. CNL targets the

Warburg effect and also mitochondrial respiration by targeting GAPDH,

suppressing STAT3 transcriptional activity, suppressing mitochondrial STAT3

function and also perturbing mitochondrial integrity by increased MOMP. This

mechanism of CNL-induced cell death in CLL cells is promising clinically

because recent reports have shown that mitochondrial respiration and stromal-

mediated glycolytic switch are robust therapeutic targets in CLL [131, 233].

Fig. 5-1 represents our proposed mechanisms of CNL-induced cell death

in CLL cells.

136

Figure 5-1. Molecular mechanisms of CNL-induced cell death in CLL cells

Future Directions

Future directions of this research include:

1. Validate the effectiveness of CNL treatment in a leukemic mouse model of

CLL like the recently described mouse model for CLL generated by co-

injection of primary human CLL cells and T cells into NOD/ SCID mice [182].

2. Evaluate the effect of CNL treatment on other components of the glycolytic

metabolism like other glycolytic enzymes and nutrient transporters.

3. Determine other potential mechanisms of CNL-mediated regulation of

GAPDH. Potential mechanisms to be evaluated include regulation at the

transcriptional, translational, post-translational modification level or enzymatic

regulation. CNL-mediated regulation of nuclear GAPDH should also be

assessed.

137

4. Metabolomic studies should be undertaken to obtain a clear picture of CNL-

induced changes in cellular metabolism. This might also shed light on the

effect of CNL treatment on cellular bioenergetics.

5. Determine which of the two phosphorylation sites i.e. S727 or Y705 is

essential for CNL-induced cell death in CLL cells.

6. Elucidate exact mechanism of CNL-induced reduction in mitochondrial

STAT3 levels in CLL cells. Assess the effect of CNL on GRIM-19 and identify

mitochondrial enzymes regulated by CNL that suppress mitochondrial STAT3

phosphorylation.

7. Elucidate the mechanism of CNL-induced suppression in oxidative

phosphorylation in CLL cells and determine if reduction in mitochondrial

p-STAT3-S727 plays a role in reduced oxidative phosphorylation.

8. Assess if CNL affects the autophagic function of cytoplasmic STAT3 and if

this mechanism plays a role in cell death.

9. Investigate if co-treatment of CNL with other inhibitors and/or mimetics of the

sphingolipid pathway enhances this induction of cell death in CLL. Utilization

of a GCS inhibitor (PPMP, PDMP, 4-HPR, etc.), ceramidase inhibitor (NOE,

LCL204, LCL385), or sphingosine kinase inhibitor (etoposide) could

potentially increase endogenous and exogenous ceramide levels and

augment cell death induction in CLL.

10. Investigate if co-treatment of CNL with STAT3 inhibitors enhances the

effectiveness of this therapy.

138

11. Develop targeted delivery of CNL by conjugating the liposome with antibody

against CLL-specific antigens like CD20 and CD74. Rituximab can be used

for targeting CD20 and mitatuzumab can be used for targeting CD74.

Summary and therapeutic potential

The role of CNL as a therapy for tumor regression in several types of solid

and non-solid cancers is extensively studied. However, the effect of CNL

treatment in CLL remains unclear. We have identified key molecular signaling

pathways of CNL-induced cell death in CLL. Identifying the key signaling

pathways inducing cell death following CNL treatment provides insight into the

therapeutic usefulness of this agent and is also valuable to uncover additional

targets for potential combination therapies.

In this dissertation we have provided convincing data demonstrating that

CNL treatment is a multi-targeted treatment strategy. CNL-induced cytotoxicity in

CLL cells is a result of targeting multiple cellular systems including cellular

bioenergetics, mitochondrial integrity and cellular protein expression profile.

Encapsulation of chemotherapeutic drugs into nanoliposomes has been a largely

successful delivery formulation in cancer models in vitro and in vivo. Extensive in

vivo toxicology studies conducted by Nanotechnology Characterization

Laboratory (http://ncl.cancer.gov/working_technical_reports.asp) have confirmed

that CNL have minimum adverse toxicity [113]. Moreover, the ongoing success of

using CNL as a platform for combinatorial therapy with other neoplastic agents

presents a promising and lucrative future to this endeavor of developing more

effective sphingolipid-based therapeutics for CLL.

139

References

1. Merrill, A.H., Jr. and G.M. Carman, Introduction to Thematic Minireview Series: Novel Bioactive Sphingolipids. J Biol Chem, 2015. 290(25): p. 15362-4.

2. Hannun, Y.A. and L.M. Obeid, Principles of bioactive lipid signalling: lessons from sphingolipids. Nat Rev Mol Cell Biol, 2008. 9(2): p. 139-50.

3. Beckham, T.H., et al., Interdiction of sphingolipid metabolism to improve standard cancer therapies. Adv Cancer Res, 2013. 117: p. 1-36.

4. Ogretmen, B. and Y.A. Hannun, Biologically active sphingolipids in cancer pathogenesis and treatment. Nat Rev Cancer, 2004. 4(8): p. 604-16.

5. Brown, D.A. and E. London, Structure and function of sphingolipid- and cholesterol-rich membrane rafts. J Biol Chem, 2000. 275(23): p. 17221-4.

6. Kolesnick, R.N., F.M. Goni, and A. Alonso, Compartmentalization of ceramide signaling: physical foundations and biological effects. J Cell Physiol, 2000. 184(3): p. 285-300.

7. Holthuis, J.C., et al., The organizing potential of sphingolipids in intracellular membrane transport. Physiol Rev, 2001. 81(4): p. 1689-723.

8. Singh, P., Y.D. Paila, and A. Chattopadhyay, Role of glycosphingolipids in the function of human serotonin(1)A receptors. J Neurochem, 2012. 123(5): p. 716-24.

9. Simons, K. and D. Toomre, Lipid rafts and signal transduction. Nat Rev Mol Cell Biol, 2000. 1(1): p. 31-9.

10. Hait, N.C., et al., Regulation of histone acetylation in the nucleus by sphingosine-1-phosphate. Science, 2009. 325(5945): p. 1254-7.

11. Zeidan, Y.H. and Y.A. Hannun, Translational aspects of sphingolipid metabolism. Trends Mol Med, 2007. 13(8): p. 327-36.

12. Ryland, L.K., et al., Dysregulation of sphingolipid metabolism in cancer. Cancer Biol Ther, 2011. 11(2): p. 138-49.

13. Obeid, L.M., et al., Programmed cell death induced by ceramide. Science, 1993. 259(5102): p. 1769-71.

14. Jarvis, W.D., et al., Induction of apoptotic DNA damage and cell death by activation of the sphingomyelin pathway. Proc Natl Acad Sci U S A, 1994. 91(1): p. 73-7.

15. Haimovitz-Friedman, A., et al., Ionizing radiation acts on cellular membranes to generate ceramide and initiate apoptosis. J Exp Med, 1994. 180(2): p. 525-35.

16. Cifone, M.G., et al., Apoptotic signaling through CD95 (Fas/Apo-1) activates an acidic sphingomyelinase. J Exp Med, 1994. 180(4): p. 1547-52.

17. Bose, R., et al., Ceramide synthase mediates daunorubicin-induced apoptosis: an alternative mechanism for generating death signals. Cell, 1995. 82(3): p. 405-14.

18. Morad, S.A. and M.C. Cabot, Ceramide-orchestrated signalling in cancer cells. Nat Rev Cancer, 2013. 13(1): p. 51-65.

19. Dumitru, C.A. and E. Gulbins, TRAIL activates acid sphingomyelinase via a redox mechanism and releases ceramide to trigger apoptosis. Oncogene, 2006. 25(41): p. 5612-25.

20. Miyaji, M., et al., Role of membrane sphingomyelin and ceramide in platform formation for Fas-mediated apoptosis. J Exp Med, 2005. 202(2): p. 249-59.

21. Grassme, H., et al., CD95 signaling via ceramide-rich membrane rafts. J Biol Chem, 2001. 276(23): p. 20589-96.

140

22. Huang, S.T., et al., Phyllanthus urinaria induces the Fas receptor/ligand expression and ceramide-mediated apoptosis in HL-60 cells. Life Sci, 2004. 75(3): p. 339-51.

23. Schaefer, J.T., W. Barthlen, and P. Schweizer, Ceramide induces apoptosis in neuroblastoma cell cultures resistant to CD95 (Fas/APO-1)-mediated apoptosis. J Pediatr Surg, 2000. 35(3): p. 473-9.

24. White-Gilbertson, S., et al., Ceramide synthase 6 modulates TRAIL sensitivity and nuclear translocation of active caspase-3 in colon cancer cells. Oncogene, 2009. 28(8): p. 1132-41.

25. Voelkel-Johnson, C., Y.A. Hannun, and A. El-Zawahry, Resistance to TRAIL is associated with defects in ceramide signaling that can be overcome by exogenous C6-ceramide without requiring down-regulation of cellular FLICE inhibitory protein. Mol Cancer Ther, 2005. 4(9): p. 1320-7.

26. Siskind, L.J. and M. Colombini, The lipids C2- and C16-ceramide form large stable channels. Implications for apoptosis. J Biol Chem, 2000. 275(49): p. 38640-4.

27. Siskind, L.J., Mitochondrial ceramide and the induction of apoptosis. J Bioenerg Biomembr, 2005. 37(3): p. 143-53.

28. Ganesan, V., et al., Ceramide and activated Bax act synergistically to permeabilize the mitochondrial outer membrane. Apoptosis, 2010. 15(5): p. 553-62.

29. von Haefen, C., et al., Ceramide induces mitochondrial activation and apoptosis via a Bax-dependent pathway in human carcinoma cells. Oncogene, 2002. 21(25): p. 4009-19.

30. Darios, F., et al., Ceramide increases mitochondrial free calcium levels via caspase 8 and Bid: role in initiation of cell death. J Neurochem, 2003. 84(4): p. 643-54.

31. Yuan, H., et al., Cytochrome c dissociation and release from mitochondria by truncated Bid and ceramide. Mitochondrion, 2003. 2(4): p. 237-44.

32. Chen, C.L., et al., Ceramide induces p38 MAPK and JNK activation through a mechanism involving a thioredoxin-interacting protein-mediated pathway. Blood, 2008. 111(8): p. 4365-74.

33. Kim, S.S., et al., P53 mediates ceramide-induced apoptosis in SKN-SH cells. Oncogene, 2002. 21(13): p. 2020-8.

34. Deng, X., F. Gao, and W.S. May, Protein phosphatase 2A inactivates Bcl2's antiapoptotic function by dephosphorylation and up-regulation of Bcl2-p53 binding. Blood, 2009. 113(2): p. 422-8.

35. Liu, X., et al., Targeting of survivin by nanoliposomal ceramide induces complete remission in a rat model of NK-LGL leukemia. Blood, 2010. 116(20): p. 4192-201.

36. Hetz, C.A., et al., Caspase-dependent initiation of apoptosis and necrosis by the Fas receptor in lymphoid cells: onset of necrosis is associated with delayed ceramide increase. J Cell Sci, 2002. 115(Pt 23): p. 4671-83.

37. Villena, J., et al., Ceramide-induced formation of ROS and ATP depletion trigger necrosis in lymphoid cells. Free Radic Biol Med, 2008. 44(6): p. 1146-60.

38. Davis, M.A., et al., Effect of ceramide on intracellular glutathione determines apoptotic or necrotic cell death of JB6 tumor cells. Toxicol Sci, 2000. 53(1): p. 48-55.

39. Ryland, L.K., et al., C6-ceramide nanoliposomes target the Warburg effect in chronic lymphocytic leukemia. PLoS One, 2013. 8(12): p. e84648.

40. Kim, W.H., et al., Ceramide induces non-apoptotic cell death in human glioma cells. Neurochem Res, 2005. 30(8): p. 969-79.

141

41. Ramos, B., et al., Prevalence of necrosis in C2-ceramide-induced cytotoxicity in NB16 neuroblastoma cells. Mol Pharmacol, 2003. 64(2): p. 502-11.

42. Gentil, B., F. Grimot, and C. Riva, Commitment to apoptosis by ceramides depends on mitochondrial respiratory function, cytochrome c release and caspase-3 activation in Hep-G2 cells. Mol Cell Biochem, 2003. 254(1-2): p. 203-10.

43. Morad, S.A. and M.C. Cabot, Ceramide-orchestrated signalling in cancer cells. Nature reviews. Cancer, 2013. 13(1): p. 51-65.

44. Grosch, S., S. Schiffmann, and G. Geisslinger, Chain length-specific properties of ceramides. Prog Lipid Res, 2012. 51(1): p. 50-62.

45. Zheng, W., et al., Ceramides and other bioactive sphingolipid backbones in health and disease: lipidomic analysis, metabolism and roles in membrane structure, dynamics, signaling and autophagy. Biochimica et biophysica acta, 2006. 1758(12): p. 1864-84.

46. Levy, M. and A.H. Futerman, Mammalian ceramide synthases. IUBMB life, 2010. 62(5): p. 347-56.

47. Brunton, L.L.L., J.S.; Parker, K.L., eds., Goodman and Gilman's The Pharmacological Basis of Therapeutics. Eleventh edition ed. 2006, New York: McGraw-Hill. 2021.

48. Bassoy, E.Y. and Y. Baran, Bioactive sphingolipids in docetaxel-induced apoptosis in human prostate cancer cells. Biomedicine & pharmacotherapy = Biomedecine & pharmacotherapie, 2012. 66(2): p. 103-10.

49. Kolesnick, R., D. Altieri, and Z. Fuks, A CERTain role for ceramide in taxane-induced cell death. Cancer Cell, 2007. 11(6): p. 473-5.

50. Charles, A.G., et al., Taxol-induced ceramide generation and apoptosis in human breast cancer cells. Cancer chemotherapy and pharmacology, 2001. 47(5): p. 444-50.

51. Di Bartolomeo, S., et al., Apoptosis induced by doxorubicin in neurotumor cells is divorced from drug effects on ceramide accumulation and may involve cell cycle-dependent caspase activation. Journal of neurochemistry, 2000. 75(2): p. 532-9.

52. Senkal, C.E., et al., Role of human longevity assurance gene 1 and C18-ceramide in chemotherapy-induced cell death in human head and neck squamous cell carcinomas. Molecular cancer therapeutics, 2007. 6(2): p. 712-22.

53. Rath, G., et al., De novo ceramide synthesis is responsible for the anti-tumor properties of camptothecin and doxorubicin in follicular thyroid carcinoma. Int J Biochem Cell Biol, 2009. 41(5): p. 1165-72.

54. Park, M.A., et al., Vorinostat and sorafenib increase CD95 activation in gastrointestinal tumor cells through a Ca(2+)-de novo ceramide-PP2A-reactive oxygen species-dependent signaling pathway. Cancer research, 2010. 70(15): p. 6313-24.

55. Biswal, S.S., et al., Changes in ceramide and sphingomyelin following fludarabine treatment of human chronic B-cell leukemia cells. Toxicology, 2000. 154(1-3): p. 45-53.

56. Perry, D.K., et al., Serine palmitoyltransferase regulates de novo ceramide generation during etoposide-induced apoptosis. The Journal of biological chemistry, 2000. 275(12): p. 9078-84.

57. Schiffmann, S., et al., Activation of ceramide synthase 6 by celecoxib leads to a selective induction of C16:0-ceramide. Biochemical pharmacology, 2010. 80(11): p. 1632-40.

58. Cabot, M.C., et al., SDZ PSC 833, the cyclosporine A analogue and multidrug resistance modulator, activates ceramide synthesis and increases vinblastine

142

sensitivity in drug-sensitive and drug-resistant cancer cells. Cancer research, 1999. 59(4): p. 880-5.

59. Lucci, A., et al., Multidrug resistance modulators and doxorubicin synergize to elevate ceramide levels and elicit apoptosis in drug-resistant cancer cells. Cancer, 1999. 86(2): p. 300-11.

60. Wang, H., A.E. Giuliano, and M.C. Cabot, Enhanced de novo ceramide generation through activation of serine palmitoyltransferase by the P-glycoprotein antagonist SDZ PSC 833 in breast cancer cells. Molecular cancer therapeutics, 2002. 1(9): p. 719-26.

61. Kawatani, M., et al., Involvement of protein kinase C-regulated ceramide generation in inostamycin-induced apoptosis. Experimental cell research, 2000. 259(2): p. 389-97.

62. Sanchez, A.M., et al., Spisulosine (ES-285) induces prostate tumor PC-3 and LNCaP cell death by de novo synthesis of ceramide and PKCzeta activation. European journal of pharmacology, 2008. 584(2-3): p. 237-45.

63. Ramer, R., et al., Ceramide is involved in r(+)-methanandamide-induced cyclooxygenase-2 expression in human neuroglioma cells. Molecular pharmacology, 2003. 64(5): p. 1189-98.

64. Gustafsson, K., et al., Potentiation of cannabinoid-induced cytotoxicity in mantle cell lymphoma through modulation of ceramide metabolism. Molecular cancer research : MCR, 2009. 7(7): p. 1086-98.

65. Ponzoni, M., et al., Differential effects of N-(4-hydroxyphenyl)retinamide and retinoic acid on neuroblastoma cells: apoptosis versus differentiation. Cancer research, 1995. 55(4): p. 853-61.

66. Reynolds, C.P., B.J. Maurer, and R.N. Kolesnick, Ceramide synthesis and metabolism as a target for cancer therapy. Cancer Lett, 2004. 206(2): p. 169-80.

67. Jiang, L., et al., Preferential involvement of both ROS and ceramide in fenretinide-induced apoptosis of HL60 rather than NB4 and U937 cells. Biochemical and biophysical research communications, 2011. 405(2): p. 314-8.

68. Wang, H., et al., N-(4-hydroxyphenyl)retinamide elevates ceramide in neuroblastoma cell lines by coordinate activation of serine palmitoyltransferase and ceramide synthase. Cancer research, 2001. 61(13): p. 5102-5.

69. Maurer, B.J., et al., Synergistic cytotoxicity in solid tumor cell lines between N-(4-hydroxyphenyl)retinamide and modulators of ceramide metabolism. Journal of the National Cancer Institute, 2000. 92(23): p. 1897-909.

70. Rahmaniyan, M., et al., Identification of dihydroceramide desaturase as a direct in vitro target for fenretinide. The Journal of biological chemistry, 2011. 286(28): p. 24754-64.

71. Kraveka, J.M., et al., Involvement of dihydroceramide desaturase in cell cycle progression in human neuroblastoma cells. The Journal of biological chemistry, 2007. 282(23): p. 16718-28.

72. Apraiz, A., et al., Evaluation of bioactive sphingolipids in 4-HPR-resistant leukemia cells. BMC Cancer, 2011. 11: p. 477.

73. Fabrias, G., et al., Dihydroceramide desaturase and dihydrosphingolipids: debutant players in the sphingolipid arena. Prog Lipid Res, 2012. 51(2): p. 82-94.

74. Jin, J., et al., AMPK inhibitor Compound C stimulates ceramide production and promotes Bax redistribution and apoptosis in MCF7 breast carcinoma cells. Journal of lipid research, 2009. 50(12): p. 2389-97.

75. Jaffrezou, J.P., et al., Daunorubicin-induced apoptosis: triggering of ceramide generation through sphingomyelin hydrolysis. The EMBO journal, 1996. 15(10): p. 2417-24.

143

76. Ito, H., et al., Transcriptional regulation of neutral sphingomyelinase 2 gene expression of a human breast cancer cell line, MCF-7, induced by the anti-cancer drug, daunorubicin. Biochimica et biophysica acta, 2009. 1789(11-12): p. 681-90.

77. Modrak, D.E., et al., Synergistic interaction between sphingomyelin and gemcitabine potentiates ceramide-mediated apoptosis in pancreatic cancer. Cancer research, 2004. 64(22): p. 8405-10.

78. Dumitru, C.A., et al., Lysosomal ceramide mediates gemcitabine-induced death of glioma cells. Journal of molecular medicine, 2009. 87(11): p. 1123-32.

79. Strum, J.C., et al., 1-beta-D-Arabinofuranosylcytosine stimulates ceramide and diglyceride formation in HL-60 cells. The Journal of biological chemistry, 1994. 269(22): p. 15493-7.

80. Noda, S., et al., Role of ceramide during cisplatin-induced apoptosis in C6 glioma cells. Journal of neuro-oncology, 2001. 52(1): p. 11-21.

81. Zeidan, Y.H., R.W. Jenkins, and Y.A. Hannun, Remodeling of cellular cytoskeleton by the acid sphingomyelinase/ceramide pathway. J Cell Biol, 2008. 181(2): p. 335-50.

82. Bezombes, C., et al., Rituximab antiproliferative effect in B-lymphoma cells is associated with acid-sphingomyelinase activation in raft microdomains. Blood, 2004. 104(4): p. 1166-73.

83. Yun, S.H., et al., Stichoposide C induces apoptosis through the generation of ceramide in leukemia and colorectal cancer cells and shows in vivo antitumor activity. Clinical cancer research : an official journal of the American Association for Cancer Research, 2012. 18(21): p. 5934-48.

84. Rubio, S., et al., Betuletol 3-methyl ether induces G(2)-M phase arrest and activates the sphingomyelin and MAPK pathways in human leukemia cells. Molecular Carcinogenesis, 2010. 49(1): p. 32-43.

85. Mondal, S., et al., Withanolide D induces apoptosis in leukemia by targeting the activation of neutral sphingomyelinase-ceramide cascade mediated by synergistic activation of c-Jun N-terminal kinase and p38 mitogen-activated protein kinase. Molecular cancer, 2010. 9: p. 239.

86. Ogretmen, B. and Y.A. Hannun, Biologically active sphingolipids in cancer pathogenesis and treatment. Nature reviews. Cancer, 2004. 4(8): p. 604-16.

87. Macchia, M., et al., Design, synthesis, and characterization of the antitumor activity of novel ceramide analogues. J Med Chem, 2001. 44(23): p. 3994-4000.

88. Liu, J., et al., Novel anti-viability ceramide analogs: design, synthesis, and structure-activity relationship studies of substituted (S)-2-(benzylideneamino)-3-hydroxy-N-tetradecylpropanamides. Bioorg Med Chem, 2010. 18(14): p. 5316-22.

89. Bieberich, E., T. Kawaguchi, and R.K. Yu, N-acylated serinol is a novel ceramide mimic inducing apoptosis in neuroblastoma cells. The Journal of biological chemistry, 2000. 275(1): p. 177-81.

90. Bieberich, E., et al., Selective apoptosis of pluripotent mouse and human stem cells by novel ceramide analogues prevents teratoma formation and enriches for neural precursors in ES cell-derived neural transplants. J Cell Biol, 2004. 167(4): p. 723-34.

91. Struckhoff, A.P., et al., Novel ceramide analogs as potential chemotherapeutic agents in breast cancer. J Pharmacol Exp Ther, 2004. 309(2): p. 523-32.

92. Novgorodov, S.A., et al., Positively charged ceramide is a potent inducer of mitochondrial permeabilization. The Journal of biological chemistry, 2005. 280(16): p. 16096-105.

144

93. Senkal, C.E., et al., Potent antitumor activity of a novel cationic pyridinium-ceramide alone or in combination with gemcitabine against human head and neck squamous cell carcinomas in vitro and in vivo. J Pharmacol Exp Ther, 2006. 317(3): p. 1188-99.

94. Hou, Q., et al., Mitochondrially targeted ceramides preferentially promote autophagy, retard cell growth, and induce apoptosis. Journal of lipid research, 2011. 52(2): p. 278-88.

95. Rossi, M.J., et al., Inhibition of growth and telomerase activity by novel cationic ceramide analogs with high solubility in human head and neck squamous cell carcinoma cells. Otolaryngol Head Neck Surg, 2005. 132(1): p. 55-62.

96. Crawford, K.W., et al., Novel ceramide analogues display selective cytotoxicity in drug-resistant breast tumor cell lines compared to normal breast epithelial cells. Cell Mol Biol (Noisy-le-grand), 2003. 49(7): p. 1017-23.

97. Adan-Gokbulut, A., et al., Novel agents targeting bioactive sphingolipids for the treatment of cancer. Curr Med Chem, 2013. 20(1): p. 108-22.

98. Barth, B.M., M.C. Cabot, and M. Kester, Ceramide-based therapeutics for the treatment of cancer. Anticancer Agents Med Chem, 2011. 11(9): p. 911-9.

99. Ganta, S., et al., Development of EGFR-Targeted Nanoemulsion for Imaging and Novel Platinum Therapy of Ovarian Cancer. Pharm Res, 2014.

100. Desai, A., T. Vyas, and M. Amiji, Cytotoxicity and apoptosis enhancement in brain tumor cells upon coadministration of paclitaxel and ceramide in nanoemulsion formulations. J Pharm Sci, 2008. 97(7): p. 2745-56.

101. Stover, T.C., et al., Thermoresponsive and biodegradable linear-dendritic nanoparticles for targeted and sustained release of a pro-apoptotic drug. Biomaterials, 2008. 29(3): p. 359-69.

102. Devalapally, H., et al., Modulation of drug resistance in ovarian adenocarcinoma by enhancing intracellular ceramide using tamoxifen-loaded biodegradable polymeric nanoparticles. Clinical cancer research : an official journal of the American Association for Cancer Research, 2008. 14(10): p. 3193-203.

103. van Vlerken, L.E., et al., Modulation of intracellular ceramide using polymeric nanoparticles to overcome multidrug resistance in cancer. Cancer research, 2007. 67(10): p. 4843-50.

104. Devalapally, H., et al., Paclitaxel and ceramide co-administration in biodegradable polymeric nanoparticulate delivery system to overcome drug resistance in ovarian cancer. Int J Cancer, 2007. 121(8): p. 1830-8.

105. Kester, M., et al., Calcium phosphate nanocomposite particles for in vitro imaging and encapsulated chemotherapeutic drug delivery to cancer cells. Nano Lett, 2008. 8(12): p. 4116-21.

106. Stover, T. and M. Kester, Liposomal delivery enhances short-chain ceramide-induced apoptosis of breast cancer cells. J Pharmacol Exp Ther, 2003. 307(2): p. 468-75.

107. Stover, T.C., et al., Systemic delivery of liposomal short-chain ceramide limits solid tumor growth in murine models of breast adenocarcinoma. Clinical cancer research : an official journal of the American Association for Cancer Research, 2005. 11(9): p. 3465-74.

108. Shabbits, J.A. and L.D. Mayer, High ceramide content liposomes with in vivo antitumor activity. Anticancer Res, 2003. 23(5A): p. 3663-9.

109. Tran, M.A., et al., Combining nanoliposomal ceramide with sorafenib synergistically inhibits melanoma and breast cancer cell survival to decrease tumor development. Clinical cancer research : an official journal of the American Association for Cancer Research, 2008. 14(11): p. 3571-81.

145

110. Tagaram, H.R., et al., Nanoliposomal ceramide prevents in vivo growth of hepatocellular carcinoma. Gut, 2011. 60(5): p. 695-701.

111. Watters, R.J., et al., Targeting glucosylceramide synthase synergizes with C6-ceramide nanoliposomes to induce apoptosis in natural killer cell leukemia. Leuk Lymphoma, 2013. 54(6): p. 1288-96.

112. Jiang, Y., et al., Combinatorial therapies improve the therapeutic efficacy of nanoliposomal ceramide for pancreatic cancer. Cancer Biol Ther, 2011. 12(7): p. 574-85.

113. Hankins, J.L., et al., The therapeutic potential of nanoscale sphingolipid technologies. Handbook of experimental pharmacology, 2013(215): p. 197-210.

114. Morad, S.A., et al., Ceramide--antiestrogen nanoliposomal combinations--novel impact of hormonal therapy in hormone-insensitive breast cancer. Molecular cancer therapeutics, 2012. 11(11): p. 2352-61.

115. Adiseshaiah, P.P., et al., Synergistic combination therapy with nanoliposomal C6-ceramide and vinblastine is associated with autophagy dysfunction in hepatocarcinoma and colorectal cancer models. Cancer letters, 2013. 337(2): p. 254-65.

116. Chapman, J.V., V. Gouaze-Andersson, and M.C. Cabot, Expression of P-glycoprotein in HeLa cells confers resistance to ceramide cytotoxicity. Int J Oncol, 2010. 37(6): p. 1591-7.

117. Cabot, M.C., T.Y. Han, and A.E. Giuliano, The multidrug resistance modulator SDZ PSC 833 is a potent activator of cellular ceramide formation. FEBS Lett, 1998. 431(2): p. 185-8.

118. Lavie, Y., et al., Agents that reverse multidrug resistance, tamoxifen, verapamil, and cyclosporin A, block glycosphingolipid metabolism by inhibiting ceramide glycosylation in human cancer cells. The Journal of biological chemistry, 1997. 272(3): p. 1682-7.

119. Morad, S.A., et al., Tamoxifen magnifies therapeutic impact of ceramide in human colorectal cancer cells independent of p53. Biochemical pharmacology, 2013. 85(8): p. 1057-65.

120. Chapman, J.V., et al., P-glycoprotein antagonists confer synergistic sensitivity to short-chain ceramide in human multidrug-resistant cancer cells. Experimental cell research, 2011. 317(12): p. 1736-45.

121. Qiu, L., et al., Paclitaxel and ceramide synergistically induce cell death with transient activation of EGFR and ERK pathway in pancreatic cancer cells. Oncol Rep, 2006. 16(4): p. 907-13.

122. Best, C., et al., Paclitaxel disrupts polarized entry of membrane-permeable C6 ceramide into ovarian cancer cells resulting in synchronous induction of cell death. Oncol Lett, 2013. 5(6): p. 1854-1858.

123. Zhu, Q.Y., et al., C6-ceramide synergistically potentiates the anti-tumor effects of histone deacetylase inhibitors via AKT dephosphorylation and alpha-tubulin hyperacetylation both in vitro and in vivo. Cell Death Dis, 2011. 2: p. e117.

124. Furlong, S.J., J.S. Mader, and D.W. Hoskin, Lactoferricin-induced apoptosis in estrogen-nonresponsive MDA-MB-435 breast cancer cells is enhanced by C6 ceramide or tamoxifen. Oncol Rep, 2006. 15(5): p. 1385-90.

125. Zenz, T., et al., From pathogenesis to treatment of chronic lymphocytic leukaemia. Nat Rev Cancer, 2010. 10(1): p. 37-50.

126. Hayden, R.E., et al., Treatment of chronic lymphocytic leukemia requires targeting of the protective lymph node environment with novel therapeutic approaches. Leuk Lymphoma, 2012. 53(4): p. 537-49.

146

127. Hanlon, K., C.E. Rudin, and L.W. Harries, Investigating the targets of MIR-15a and MIR-16-1 in patients with chronic lymphocytic leukemia (CLL). PLoS One, 2009. 4(9): p. e7169.

128. Chavez, J.C., et al., Genomic aberrations deletion 11q and deletion 17p independently predict for worse progression-free and overall survival after allogeneic hematopoietic cell transplantation for chronic lymphocytic leukemia. Leuk Res, 2014. 38(10): p. 1165-72.

129. Dohner, H., et al., Genomic aberrations and survival in chronic lymphocytic leukemia. N Engl J Med, 2000. 343(26): p. 1910-6.

130. Ghia, P., et al., Differential effects on CLL cell survival exerted by different microenvironmental elements. Curr Top Microbiol Immunol, 2005. 294: p. 135-45.

131. Jitschin, R., et al., Stromal cell-mediated glycolytic switch in CLL cells involves Notch-c-Myc signaling. Blood, 2015. 125(22): p. 3432-6.

132. Tam, C.S., et al., Long-term results of the fludarabine, cyclophosphamide, and rituximab regimen as initial therapy of chronic lymphocytic leukemia. Blood, 2008. 112(4): p. 975-80.

133. Hallek, M., et al., Addition of rituximab to fludarabine and cyclophosphamide in patients with chronic lymphocytic leukaemia: a randomised, open-label, phase 3 trial. Lancet, 2010. 376(9747): p. 1164-74.

134. Smolej, L., Targeted treatment for chronic lymphocytic leukemia: clinical potential of obinutuzumab. Pharmgenomics Pers Med, 2015. 8: p. 1-7.

135. Shanafelt, T., Treatment of older patients with chronic lymphocytic leukemia: key questions and current answers. Hematology Am Soc Hematol Educ Program, 2013. 2013: p. 158-67.

136. Sutton, L.A. and R. Rosenquist, Deciphering the molecular landscape in chronic lymphocytic leukemia: time frame of disease evolution. Haematologica, 2015. 100(1): p. 7-16.

137. Mengubas, K., et al., Ceramide-induced killing of normal and malignant human lymphocytes is by a non-apoptotic mechanism. Oncogene, 1999. 18(15): p. 2499-506.

138. Hammadi, M., et al., Membrane microdomain sphingolipids are required for anti-CD20-induced death of chronic lymphocytic leukemia B cells. Haematologica, 2012. 97(2): p. 288-96.

139. Schwamb, J., et al., B-cell receptor triggers drug sensitivity of primary CLL cells by controlling glucosylation of ceramides. Blood, 2012. 120(19): p. 3978-85.

140. Trojani, A., et al., Gene expression profiling identifies ARSD as a new marker of disease progression and the sphingolipid metabolism as a potential novel metabolism in chronic lymphocytic leukemia. Cancer Biomark, 2011. 11(1): p. 15-28.

141. Kolesnick, R.N. and M. Kronke, Regulation of ceramide production and apoptosis. Annu Rev Physiol, 1998. 60: p. 643-65.

142. Daido, S., et al., Pivotal role of the cell death factor BNIP3 in ceramide-induced autophagic cell death in malignant glioma cells. Cancer Res, 2004. 64(12): p. 4286-93.

143. Chalfant, C.E., et al., FAS activation induces dephosphorylation of SR proteins; dependence on the de novo generation of ceramide and activation of protein phosphatase 1. J Biol Chem, 2001. 276(48): p. 44848-55.

144. Liu, X., et al., Targeting of survivin by nanoliposomal ceramide induces complete remission in NK-LGL leukemia. Blood, 2010. 116(20): p. 4192-201.

147

145. Clarke, C.J., et al., The extended family of neutral sphingomyelinases. Biochemistry, 2006. 45(38): p. 11247-56.

146. Fox, T.E., et al., Ceramide recruits and activates protein kinase C zeta (PKC zeta) within structured membrane microdomains. J Biol Chem, 2007. 282(17): p. 12450-7.

147. Tran, M.A., et al., Combining nanoliposomal ceramide with sorafenib synergistically inhibits melanoma and breast cancer cell survival to decrease tumor development. Clin Cancer Res, 2008. 14(11): p. 3571-81.

148. Guenther, G.G., et al., Ceramide starves cells to death by downregulating nutrient transporter proteins. Proc Natl Acad Sci U S A, 2008. 105(45): p. 17402-7.

149. Vander Heiden, M.G., L.C. Cantley, and C.B. Thompson, Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science, 2009. 324(5930): p. 1029-33.

150. Warburg, O., On the origin of cancer cells. Science, 1956. 123(3191): p. 309-14. 151. Altenberg, B. and K.O. Greulich, Genes of glycolysis are ubiquitously

overexpressed in 24 cancer classes. Genomics, 2004. 84(6): p. 1014-20. 152. Shanmugam, M., S.K. McBrayer, and S.T. Rosen, Targeting the Warburg effect

in hematological malignancies: from PET to therapy. Curr Opin Oncol., 2009. 21(6): p. 531-6.

153. Yeluri, S., et al., Cancer's craving for sugar:an opportunity for clinical exploitation. J Cancer Res Clin Oncol, 2009. 135(7): p. 867-77.

154. Watson, D.G., et al., The roles of sphingosine kinases 1 and 2 in regulating the Warburg effect in prostate cancer cells. Cell Signal, 2013. 25(4): p. 1011-7.

155. Foon, K.A., K.R. Rai, and R.P. Gale, Chronic lymphocytic leukemia: new insights into biology and therapy. Ann Intern Med, 1990. 113(7): p. 525-39.

156. Hofheinz, R.D., et al., Liposomal encapsulated anti-cancer drugs. Anticancer Drugs, 2005. 16(7): p. 691-707.

157. Zhao, X., et al., Liposomal coencapsulated fludarabine and mitoxantrone for lymphoproliferative disorder treatment. J Pharm Sci, 2008. 97(4): p. 1508-18.

158. Biswal, S.S., et al., Changes in ceramide and sphingomyelin following fludarabine treatment of human chronic B-cell leukemia cells. Toxicology, 2000. 154: p. 45-53.

159. Stover, T.C., et al., Systemic delivery of liposomal short-chain ceramide limits solid tumor growth in murine models of breast adenocarcinoma. Clin Cancer Res, 2005. 11(9): p. 3465-74.

160. Yang, J., et al., Platelet-derived growth factor mediates survival of leukemic large granular lymphocytes via an autocrine regulatory pathway. Blood, 2010. 115(1): p. 51-60.

161. Sato, T., et al., FAP-1: a protein tyrosine phosphatase that associates with Fas. Science, 1995. 268(5209): p. 411-5.

162. Livak, K.J. and T.D. Schmittgen, Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods, 2001. 25(4): p. 402-8.

163. Kueck, A., et al., Resveratrol inhibits glucose metabolism in human ovarian cancer cells. Gynecol Oncol, 2007. 107(3): p. 450-7.

164. Melo, J.V., et al., Two new cell lines from B-prolymphocytic leukaemia: characterization by morphology, immunological markers, karyotype and Ig gene rearrangement. Int J Cancer, 1986. 38(4): p. 531-8.

148

165. Veldurthy, A., et al., The kinase inhibitor dasatinib induces apoptosis in chronic lymphocytic leukemia cells in vitro with preference for a subgroup of patients with unmutated IgVH genes. Blood, 2008. 112(4): p. 1443-52.

166. Rathmell, J.C. and S. Kornbluth, Filling a GAP(DH) in caspase-independent cell death. Cell, 2007. 129(5): p. 861-3.

167. Colell, A., et al., GAPDH and autophagy preserve survival after apoptotic cytochrome c release in the absence of caspase activation. Cell, 2007. 129(5): p. 983-97.

168. Song, S. and T. Finkel, GAPDH and the search for alternative energy. Nat Cell Biol, 2007. 9(8): p. 869-70.

169. Loisel, S., et al., Establishment of a novel human B-CLL-like xenograft model in nude mouse. Leuk Res, 2005. 29(11): p. 1347-52.

170. Scarlatti, F., et al., Ceramide-mediated macroautophagy involves inhibition of protein kinase B and up-regulation of beclin 1. J Biol Chem, 2004. 279(18): p. 18384-91.

171. Ogretmen, B., et al., Biochemical mechanisms of the generation of endogenous long chain ceramide in response to exogenous short chain ceramide in the A549 human lung adenocarcinoma cell line. Role for endogenous ceramide in mediating the action of exogenous ceramide. J Biol Chem, 2002. 277(15): p. 12960-9.

172. Chapman, J.V., et al., Metabolism of short-chain ceramide by human cancer cells--implications for therapeutic approaches. Biochem Pharmacol, 2010. 80(3): p. 308-15.

173. Bustin, S.A., Absolute quantification of mRNA using real-time reverse transcription polymerase chain reaction assays. J Mol Endocrinol, 2000. 25(2): p. 169-93.

174. Revillion, F., et al., Glyceraldehyde-3-phosphate dehydrogenase gene expression in human breast cancer. Eur J Cancer, 2000. 36(8): p. 1038-42.

175. Lavallard, V.J., et al., Modulation of caspase-independent cell death leads to resensitization of imatinib mesylate-resistant cells. Cancer Res, 2009. 69(7): p. 3013-20.

176. Sirover, M.A., New nuclear functions of the glycolytic protein, glyceraldehyde-3-phosphate dehydrogenase, in mammalian cells. J Cell Biochem, 2005. 95(1): p. 45-52.

177. Sundararaj, K.P., et al., Rapid shortening of telomere length in response to ceramide involves the inhibition of telomere binding activity of nuclear glyceraldehyde-3-phosphate dehydrogenase. J Biol Chem, 2004. 279(7): p. 6152-62.

178. Rivenzon-Segal, D., et al., Glucose transporters and transport kinetics in retinoic acid-differentiated T47D human breast cancer cells. Am J Physiol Endocrinol Metab, 2000. 279(3): p. E508-19.

179. Huttmann, A., et al., Gene expression signatures separate B-cell chronic lymphocytic leukaemia prognostic subgroups defined by ZAP-70 and CD38 expression status. Leukemia, 2006. 20(10): p. 1774-82.

180. Geschwind, J.F., et al., Novel therapy for liver cancer: direct intraarterial injection of a potent inhibitor of ATP production. Cancer Res, 2002. 62(14): p. 3909-13.

181. Kumagai, S., R. Narasaki, and K. Hasumi, Glucose-dependent active ATP depletion by koningic acid kills high-glycolytic cells. Biochem Biophys Res Commun, 2008. 365(2): p. 362-8.

149

182. Bagnara, D., et al., A novel adoptive transfer model of chronic lymphocytic leukemia suggests a key role for T lymphocytes in the disease. Blood, 2011. 117(20): p. 5463-72.

183. Caligaris-Cappio, F. and T.J. Hamblin, B-cell chronic lymphocytic leukemia: a bird of a different feather. J Clin Oncol, 1999. 17(1): p. 399-408.

184. Hankins, J.L., et al., The therapeutic potential of nanoscale sphingolipid technologies. Handb Exp Pharmacol, 2013(215): p. 197-210.

185. Haakenson, J.K., et al., Lysosomal Degradation of CD44 Mediates Ceramide Nanoliposome-induced Anoikis and Diminshed Extravasation in Metastatic Carcinoma Cells. J Biol Chem, 2015.

186. Lin, T.S., S. Mahajan, and D.A. Frank, STAT signaling in the pathogenesis and treatment of leukemias. Oncogene, 2000. 19(21): p. 2496-504.

187. Yu, H., et al., Revisiting STAT3 signalling in cancer: new and unexpected biological functions. Nat Rev Cancer, 2014. 14(11): p. 736-46.

188. Frank, D.A., S. Mahajan, and J. Ritz, B lymphocytes from patients with chronic lymphocytic leukemia contain signal transducer and activator of transcription (STAT) 1 and STAT3 constitutively phosphorylated on serine residues. J Clin Invest, 1997. 100(12): p. 3140-8.

189. Hazan-Halevy, I., et al., STAT3 is constitutively phosphorylated on serine 727 residues, binds DNA, and activates transcription in CLL cells. Blood, 2010. 115(14): p. 2852-63.

190. Capron, C., et al., Viability and stress protection of chronic lymphoid leukemia cells involves overactivation of mitochondrial phosphoSTAT3Ser727. Cell Death Dis, 2014. 5: p. e1451.

191. Wegrzyn, J., et al., Function of mitochondrial Stat3 in cellular respiration. Science, 2009. 323(5915): p. 793-7.

192. Li, P., et al., Signal transducer and activator of transcription-3 induces microRNA-155 expression in chronic lymphocytic leukemia. PLoS One, 2013. 8(6): p. e64678.

193. Liu, Z., et al., STAT-3 activates NF-kappaB in chronic lymphocytic leukemia cells. Mol Cancer Res, 2011. 9(4): p. 507-15.

194. Badoux, X., et al., Cross-talk between chronic lymphocytic leukemia cells and bone marrow endothelial cells: role of signal transducer and activator of transcription 3. Hum Pathol, 2011. 42(12): p. 1989-2000.

195. Giannoni, P., et al., An interaction between hepatocyte growth factor and its receptor (c-MET) prolongs the survival of chronic lymphocytic leukemic cells through STAT3 phosphorylation: a potential role of mesenchymal cells in the disease. Haematologica, 2011. 96(7): p. 1015-23.

196. Lu, K., et al., The STAT3 inhibitor WP1066 reverses the resistance of chronic lymphocytic leukemia cells to histone deacetylase inhibitors induced by interleukin-6. Cancer Lett, 2015. 359(2): p. 250-8.

197. Ishdorj, G., J.B. Johnston, and S.B. Gibson, Inhibition of constitutive activation of STAT3 by curcurbitacin-I (JSI-124) sensitized human B-leukemia cells to apoptosis. Mol Cancer Ther, 2010. 9(12): p. 3302-14.

198. Schust, J., et al., Stattic: a small-molecule inhibitor of STAT3 activation and dimerization. Chem Biol, 2006. 13(11): p. 1235-42.

199. Allen, J.C., et al., c-Abl regulates Mcl-1 gene expression in chronic lymphocytic leukemia cells. Blood, 2011. 117(8): p. 2414-22.

200. Martinez-Lostao, L., et al., Role of the STAT1 pathway in apoptosis induced by fludarabine and JAK kinase inhibitors in B-cell chronic lymphocytic leukemia. Leuk Lymphoma, 2005. 46(3): p. 435-42.

150

201. Frank, D.A., S. Mahajan, and J. Ritz, Fludarabine-induced immunosuppression is associated with inhibition of STAT1 signaling. Nat Med, 1999. 5(4): p. 444-7.

202. Battle, T.E. and D.A. Frank, STAT1 mediates differentiation of chronic lymphocytic leukemia cells in response to Bryostatin 1. Blood, 2003. 102(8): p. 3016-24.

203. Kearney, C.J., et al., Necroptosis suppresses inflammation via termination of TNF- or LPS-induced cytokine and chemokine production. Cell Death Differ, 2015.

204. Miao, B. and A. Degterev, Methods to analyze cellular necroptosis. Methods Mol Biol, 2009. 559: p. 79-93.

205. Gonzalez, D., et al., Mutational status of the TP53 gene as a predictor of response and survival in patients with chronic lymphocytic leukemia: results from the LRF CLL4 trial. J Clin Oncol, 2011. 29(16): p. 2223-9.

206. Varnai, P., K.I. Rother, and T. Balla, Phosphatidylinositol 3-kinase-dependent membrane association of the Bruton's tyrosine kinase pleckstrin homology domain visualized in single living cells. J Biol Chem, 1999. 274(16): p. 10983-9.

207. Rawlings, D.J., et al., Activation of BTK by a phosphorylation mechanism initiated by SRC family kinases. Science, 1996. 271(5250): p. 822-5.

208. Byrd, J.C., et al., Targeting BTK with ibrutinib in relapsed chronic lymphocytic leukemia. N Engl J Med, 2013. 369(1): p. 32-42.

209. Pepper, C., et al., Mcl-1 expression has in vitro and in vivo significance in chronic lymphocytic leukemia and is associated with other poor prognostic markers. Blood, 2008. 112(9): p. 3807-17.

210. Awan, F.T., et al., Mcl-1 expression predicts progression-free survival in chronic lymphocytic leukemia patients treated with pentostatin, cyclophosphamide, and rituximab. Blood, 2009. 113(3): p. 535-7.

211. Hussain, S.R., et al., Mcl-1 is a relevant therapeutic target in acute and chronic lymphoid malignancies: down-regulation enhances rituximab-mediated apoptosis and complement-dependent cytotoxicity. Clin Cancer Res, 2007. 13(7): p. 2144-50.

212. Grzybowska-Izydorczyk, O., et al., Expression and prognostic significance of the inhibitor of apoptosis protein (IAP) family and its antagonists in chronic lymphocytic leukaemia. Eur J Cancer, 2010. 46(4): p. 800-10.

213. Purroy, N., et al., Targeting the proliferative and chemoresistant compartment in chronic lymphocytic leukemia by inhibiting survivin protein. Leukemia, 2014. 28(10): p. 1993-2004.

214. Loeder, S., et al., A novel paradigm to trigger apoptosis in chronic lymphocytic leukemia. Cancer Res, 2009. 69(23): p. 8977-86.

215. Kater, A.P., et al., Inhibitors of XIAP sensitize CD40-activated chronic lymphocytic leukemia cells to CD95-mediated apoptosis. Blood, 2005. 106(5): p. 1742-8.

216. Maziere, C., M.A. Conte, and J.C. Maziere, Activation of the JAK/STAT pathway by ceramide in cultured human fibroblasts. FEBS Lett, 2001. 507(2): p. 163-8.

217. Woyach, J.A., et al., Bruton's tyrosine kinase (BTK) function is important to the development and expansion of chronic lymphocytic leukemia (CLL). Blood, 2014. 123(8): p. 1207-13.

218. Kazi, J.U., N.N. Kabir, and L. Ronnstrand, Protein kinase C (PKC) as a drug target in chronic lymphocytic leukemia. Med Oncol, 2013. 30(4): p. 757.

219. Chan, K.S., et al., Forced expression of a constitutively active form of Stat3 in mouse epidermis enhances malignant progression of skin tumors induced by two-stage carcinogenesis. Oncogene, 2008. 27(8): p. 1087-94.

151

220. Barbieri, I., et al., Constitutively active Stat3 enhances neu-mediated migration and metastasis in mammary tumors via upregulation of Cten. Cancer Res, 2010. 70(6): p. 2558-67.

221. Demaria, M., et al., A STAT3-mediated metabolic switch is involved in tumour transformation and STAT3 addiction. Aging (Albany NY), 2010. 2(11): p. 823-42.

222. Poli, V. and A. Camporeale, STAT3-Mediated Metabolic Reprograming in Cellular Transformation and Implications for Drug Resistance. Front Oncol, 2015. 5: p. 121.

223. Gough, D.J., et al., Mitochondrial STAT3 supports Ras-dependent oncogenic transformation. Science, 2009. 324(5935): p. 1713-6.

224. Gough, D.J., et al., STAT3 supports experimental K-RasG12D-induced murine myeloproliferative neoplasms dependent on serine phosphorylation. Blood, 2014. 124(14): p. 2252-61.

225. Mackenzie, G.G., et al., Targeting mitochondrial STAT3 with the novel phospho-valproic acid (MDC-1112) inhibits pancreatic cancer growth in mice. PLoS One, 2013. 8(5): p. e61532.

226. Zhang, Q., et al., Mitochondrial localized Stat3 promotes breast cancer growth via phosphorylation of serine 727. J Biol Chem, 2013. 288(43): p. 31280-8.

227. Myers, M.G., Jr., Cell biology. Moonlighting in mitochondria. Science, 2009. 323(5915): p. 723-4.

228. Garcia, R., et al., Constitutive activation of Stat3 by the Src and JAK tyrosine kinases participates in growth regulation of human breast carcinoma cells. Oncogene, 2001. 20(20): p. 2499-513.

229. Schindler, C., D.E. Levy, and T. Decker, JAK-STAT signaling: from interferons to cytokines. J Biol Chem, 2007. 282(28): p. 20059-63.

230. Gough, D.J., L. Koetz, and D.E. Levy, The MEK-ERK pathway is necessary for serine phosphorylation of mitochondrial STAT3 and Ras-mediated transformation. PLoS One, 2013. 8(11): p. e83395.

231. Tammineni, P., et al., The import of the transcription factor STAT3 into mitochondria depends on GRIM-19, a component of the electron transport chain. J Biol Chem, 2013. 288(7): p. 4723-32.

232. Capron, C., et al., Viability and stress protection of chronic lymphoid leukemia cells involves overactivation of mitochondrial phosphoSTAT3Ser727. Cell Death Dis, 2014. 5: p. e1451.

233. Jitschin, R., et al., Mitochondrial metabolism contributes to oxidative stress and reveals therapeutic targets in chronic lymphocytic leukemia. Blood, 2014. 123(17): p. 2663-72.

234. Zamzami, N., et al., Sequential reduction of mitochondrial transmembrane potential and generation of reactive oxygen species in early programmed cell death. J Exp Med, 1995. 182(2): p. 367-77.

235. Benov, L., L. Sztejnberg, and I. Fridovich, Critical evaluation of the use of hydroethidine as a measure of superoxide anion radical. Free Radic Biol Med, 1998. 25(7): p. 826-31.

236. Birbes, H., et al., Mitochondria and ceramide: intertwined roles in regulation of apoptosis. Adv Enzyme Regul, 2002. 42: p. 113-29.

237. Quillet-Mary, A., et al., Implication of mitochondrial hydrogen peroxide generation in ceramide-induced apoptosis. J Biol Chem, 1997. 272(34): p. 21388-95.

238. Scharstuhl, A., et al., Involvement of VDAC, Bax and ceramides in the efflux of AIF from mitochondria during curcumin-induced apoptosis. PLoS One, 2009. 4(8): p. e6688.

152

239. Czubowicz, K. and R. Strosznajder, Ceramide in the molecular mechanisms of neuronal cell death. The role of sphingosine-1-phosphate. Mol Neurobiol, 2014. 50(1): p. 26-37.

240. Kim, N.H., et al., PKB/Akt inhibits ceramide-induced apoptosis in neuroblastoma cells by blocking apoptosis-inducing factor (AIF) translocation. J Cell Biochem, 2007. 102(5): p. 1160-70.

241. Pardo, J., et al., A role of the mitochondrial apoptosis-inducing factor in granulysin-induced apoptosis. J Immunol, 2001. 167(3): p. 1222-9.

242. Delavallee, L., et al., AIF-mediated caspase-independent necroptosis: a new chance for targeted therapeutics. IUBMB Life, 2011. 63(4): p. 221-32.

243. Flaveny, C.A., et al., Broad Anti-tumor Activity of a Small Molecule that Selectively Targets the Warburg Effect and Lipogenesis. Cancer Cell, 2015. 28(1): p. 42-56.

244. Zhang, P., et al., C6-ceramide nanoliposome suppresses tumor metastasis by eliciting PI3K and PKCzeta tumor-suppressive activities and regulating integrin affinity modulation. Sci Rep, 2015. 5: p. 9275.

245. El-Gamal, D., et al., PKC-beta as a therapeutic target in CLL: PKC inhibitor AEB071 demonstrates preclinical activity in CLL. Blood, 2014. 124(9): p. 1481-91.

246. Stathem, M., et al., Glucose availability and glycolytic metabolism dictate glycosphingolipid levels. J Cell Biochem, 2015. 116(1): p. 67-80.

247. Kogot-Levin, A. and A. Saada, Ceramide and the mitochondrial respiratory chain. Biochimie, 2014. 100: p. 88-94.

153

Appendix: Letters of Permission

Chapter 1 and Chapter 2 has material that has been previously published. This

material is used with permission from the respective sources. Copies of the letter

granting permission for use in this dissertation are as follows.

154

Permission letter for Chapter 1

Reprinted (adapted) with permission from the chapter:

Doshi UA, Haakenson JK, Linton SL, Kelly K, Kester M (2015). Chemotherapy

and sphingolipid metabolism. In Bioactive Sphingolipids in Cancer Biology and

Therapy (pp. 401-436), New York, NY: Springer.

155

Permission for Chapter 2

Reprinted (adapted) with permission from:

Ryland LK*, Doshi UA*, Shanmugavelandy SS, Fox TE, Aliaga C, Broeg K, Baab

KT, Young M, Khan O, Haakenson JK, Jarbadan NR, Liao J, Wang HG, Feith

DJ, Loughran TP Jr, Liu X, Kester M. “C6-ceramide nanoliposomes target the

Warburg effect in chronic lymphocytic leukemia” PLoS One. 2013 Dec 19;

8(12):e84648.

According to the journal policy:

“PLOS applies the Creative Commons Attribution (CC BY) license to works we publish. Using PLOS Content

No permission is required from the authors or the publishers to reuse or

repurpose PLOS content provided the original article is cited. In most cases,

appropriate attribution can be provided by simply citing the original article.”

http://journals.plos.org/plosone/s/content-license

VITA

USHMA A. DOSHI

Education

2015 Ph.D. Molecular Medicine, Pennsylvania State University 2015 MBA Pennsylvania State University 2012 M.S. Biotechnology, University at Buffalo, New York 2008 B.S. Pharmaceutical Sc., Institute of Chemical Technology, Mumbai, India

Publications Doshi UA, Haakenson JK, Linton SL, Kelly K, Kester M (2015). Chemotherapy and

sphingolipid metabolism. In Bioactive Sphingolipids in Cancer Biology and Therapy

(pp. 401-436), New York, NY: Springer.

Ryland LK*, Doshi UA*, Shanmugavelandy SS, Fox TE, Aliaga C, Broeg K, Baab KT, Young M, Khan O, Haakenson JK, Jarbadan NR, Liao J, Wang HG, Feith DJ, Loughran TP Jr, Liu X, Kester M. C6-ceramide nanoliposomes target the Warburg effect in chronic lymphocytic leukemia. PLoS One. 2013 Dec 19;8(12):e84648. *Co-first authorship

Hankins JL, Doshi UA, Haakenson JK, Young MM, Barth BM, Kester M. The therapeutic potential of nanoscale sphingolipid technologies. Handb Exp Pharmacol. 2013;(215):197-210.

Manuscripts in preparation:

Doshi UA, Fox TE, Loughran TP, Kester M. STAT3 mediates nanoliposomal C6-ceramide-induced cell death in chronic lymphocytic leukemia.

Doshi UA, Fox TE, Loughran TP, Kester M. Nanoliposomal C6-ceramide targets cellular bioenergetics to induce cell death in chronic lymphocytic leukemia.

Awards College of Medicine Class of 1971 Endowed Scholarship, Pennsylvania State

University, 2014

Beta Gamma Sigma Inductee, The Penn State Capital College Chapter of Beta Gamma Sigma scholastic honor society in Business Administration, 2014

Outstanding Graduate Student Award, 7th International Ceramide Conference , 2013

University Graduate Fellowship, Pennsylvania State University, 2010