SlideShare a Scribd company logo
1 of 56
Yale University
EliScholar – A Digital Platform for Scholarly Publishing at Yale
Yale Medicine Thesis Digital Library School of Medicine
January 2019
Development Of Pancreatic Cancer Organoid
Model For Studying Immune Response In
Pancreatic Cancer
Jin Woo Yoo
Follow this and additional works at: https://elischolar.library.yale.edu/ymtdl
This Open Access Thesis is brought to you for free and open access by the School of Medicine at EliScholar – A Digital Platform for Scholarly
Publishing at Yale. It has been accepted for inclusion in Yale Medicine Thesis Digital Library by an authorized administrator of EliScholar – A Digital
Platform for Scholarly Publishing at Yale. For more information, please contact elischolar@yale.edu.
Recommended Citation
Yoo, Jin Woo, "Development Of Pancreatic Cancer Organoid Model For Studying Immune Response In Pancreatic Cancer" (2019).
Yale Medicine Thesis Digital Library. 3543.
https://elischolar.library.yale.edu/ymtdl/3543
Development of Pancreatic Cancer Organoid Models for
Studying Immune Response in Pancreatic Cancer
A Thesis Submitted to the
Yale University School of Medicine
in Partial Fulfillment of the Requirements for the
Degree of Doctor of Medicine
by
Jin Woo Yoo
2019
DEVELOPMENT OF PANCREATIC CANCER ORGANOID MODEL FOR STUDYING
IMMUNE RESPONSE IN PANCREATIC CANCER. Jin Woo Yoo, Prashanth R. Gokare,
Yevgeniya Foster, Brittany Fitzgerald, Nikhil S. Joshi, James J. Farrell. Section of
Gastroenterology, Department of Internal Medicine, Yale University, School of Medicine, New
Haven, CT.
The importance of immune system in pancreatic ductal adenocarcinoma (PDAC)
pathogenesis and therapy remains poorly understood largely due to the lack of effective model
systems. Cell lines are not physiologic as they cannot recapitulate the cancer stroma and lose
genetic heterogeneity over time. Genetically engineered mouse models of PDAC are more
physiologic than cell lines but lack neoantigens needed to mount T cell responses against tumor.
Organoid models of PDAC offer unique opportunity to study immune mechanisms in PDAC
since organoids can model complex layering of multiple cell types, creating a physiologically
relevant system that is highly tractable for genetic manipulation, co-cultures, and high
throughput assays. In this study, we sought to establish murine and human organoid models of
PDAC to investigate the biology of PDAC immune response, with the specific aims of
developing transplantable immunogenic murine PDAC organoid models for the study of antigen-
specific anti-tumor T cell responses and assembling a library of experimentally validated,
patient-derived PDAC organoid lines for pancreatic cancer precision medicine research.
To generate immunogenic murine organoid models of PDAC, pancreatic organoids were
isolated from “KP-NINJA” (KrasLox-STOP-Lox-G12D
; P53flox/flox
; inversion induced joined
neoantigen) mouse model that has been genetically engineered to express GFP-tagged T cell
neoantigens derived from lymphocytic choriomeningitis virus in an inducible fashion. Isolated
organoids were transformed in vitro using a lentiviral construct encoding Cre recombinase and
RFP reporter for expression of oncogenic KRAS and deletion of P53. A subset of transformed
organoids was additionally treated with an adenoviral construct encoding FLPo recombinase to
turn on neoantigen expression. Transformed organoids were combined with T cells in both in
vivo and in vitro setting to assess for impact on tumor growth. Patient-derived PDAC organoids
were generated using endoscopic ultrasound-guided fine needle biopsy (EUS-FNB) specimens,
surgical resection specimens, and tissues from patient-derived xenograft mouse models of
PDAC. Established human organoid lines were validated by Sanger sequencing, tumor formation
in vivo and immunohistochemistry of organoid-derived tumors.
Subcutaneous injection of transformed murine PDAC organoids formed tumors in mouse
that are histologically similar to early lesions found in human PDAC. Serial in vivo transfer of
these organoids by performing sequential rounds of organoid generation from tumors derived
from organoids formed progressively more advanced tumors. High level of neoantigen
expression in 100% of cells comprising murine PDAC organoids resulted in rejection of tumor
growth in mouse, while a low level of neoantigen expression restricted to 10% of cells permitted
tumor growth with increased immune infiltration. Expression of neoantigens in T cell-PDAC
organoid co-culture model systems promoted T cell infiltration of basement membrane matrix.
Additionally, we generated 30+ patient-derived PDAC organoid lines using EUS-FNB and
surgical specimens at Yale from 10/2017 to 5/2018.
We have successfully established murine and human organoid models of PDAC from
various tissues capturing discrete stages of PDAC progression. Our murine organoid models are
uniquely equipped to study antigen-specific T cell responses against tumor. Ongoing work
includes using CRISPR/Cas9-based lentiviral systems to define genes that impact anti-tumor T
cell responses and using patient-derived organoids for precision medicine research.
ACKNOWLEDGEMENTS
Work for this thesis was completed in the Joshi laboratory under the co-mentorship of
James J. Farrell, MD and Nikhil S. Joshi, PhD. Both Dr. Farrell and Dr. Joshi suggested
experiments and supervised the work done. Dr. Joshi developed the KP-NINJA mouse model
that was fundamental for the creation of immunogenic murine PDAC organoid models. Dr.
Farrell performed and provided all the endoscopic ultrasound-guided fine-needle biopsies for the
creation of patient-derived PDAC organoid lines. Prashanth Gokare, PhD collaborated with the
author on the development of three-dimensional co-culture system for murine pancreatic cancer
organoids and T cells and the creation of patient-derived PDAC organoids from surgical
resection specimens and their sequencing. Yevgeniya Foster, MD collaborated with the author on
creation of immunogenic murine PDAC organoid lines for characterizing immune responses in
vivo and immunohistochemical analysis of murine organoid-derived tumors. Brittany Fitzgerald
established the primary murine pancreatic cancer cell lines from KP-C mouse and collaborated
with the author on in vivo transfer of P14 mouse splenocytes and in vivo imaging for luciferase
detection. Marie Robert, MD provided surgical resection specimens for creation of patient-
derived pancreatic cancer organoids and interpretation of tumor histology. Ryan Sowell, PhD
from Kaech laboratory created the patient-derived xenograft mouse models, some of which were
used as source material for the creation of human PDAC organoids. All other experiments were
performed independently by the author. Dr. Farrell and Dr. Joshi reviewed and provided
comments on the manuscript. National Institute of Health-National Institute of Diabetes and
Digestive and Kidney Diseases Medical Student Research Fellowship (T35 grant), Yale
University School of Medicine Research Fellowship, and Richard Alan HirshField Memorial
Fellowship provided funding to support this work.
TABLE OF CONTENTS
LIST OF ABBREVIATIONS.......................................................................................................1
INTRODUCTION.........................................................................................................................2
• Background
• Cell of Origin
• Genetic Landscape of Pancreatic Cancer
• Precursor Lesions
• Mutational Processes
• Tumoral Heterogeneity
• Molecular Subtyping of Pancreatic Cancer
• Deranged Signaling Pathways / Molecular Aberrations
• Tumor Microenvironment
• Metabolic Reprogramming
• Immune Response in Pancreatic Cancer is Unclear
• Pre-clinical Modeling of Pancreatic Cancer
STATEMENT OF PURPOSE....................................................................................................18
METHODS...................................................................................................................................19
• Acquisition of human specimens
• Isolation and culture of murine pancreatic organoids
• Isolation and culture of human PDAC organoids
• Isolation of primary murine PDAC cell lines
• Genetic manipulation of murine pancreatic organoids
• In vivo mouse assays
• Immunohistochemical analysis of tumors
• Sanger sequencing of organoids
• Development of organoid-T cell co-culture model systems
RESULTS.....................................................................................................................................26
• KP-NINJA mouse model provides substrate for creation of immunogenic murine
organoid models of PDAC
• In vitro transformed murine pancreatic organoids form tumors that are histologically
similar to early lesions found in human PDAC
• Serial in vivo transfer of transformed murine pancreatic organoids results in
progressively more advanced tumors
• Expression of neoantigens in murine PDAC organoids elicits effective immune response
in mouse
• Expression of neoantigens in murine PDAC organoids promotes T cell infiltration in T
cell-organoid co-culture model
• Assembly of human PDAC organoid library
1
DISCUSSION...............................................................................................................................33
REFERENCES.............................................................................................................................36
FIGURES......................................................................................................................................38
TABLES........................................................................................................................................48
LIST OF ABBREVIATIONS
CTGF Connective tissue growth factor
EGF Epidermal growth factor
ER Estrogen receptor
ETC Electron transport chain
EUS-FNB Endoscopic ultrasound-guided fine needle
biopsy
FGF Fibroblast growth factor
FLP Flippase
FRT Flippase recognition target
GFP Green fluorescent protein
GM-CSF Granulocyte-macrophage colony-stimulating
factor
GP Glycoprotein
hENT1 Human equilibrative nucleoside transporter
HGF Hepatocyte growth factor
HIF1α Hypoxia-inducible transcription factor 1α
HR Homologous recombination
IFN-γ Interferon-γ
IGF1 Insulin-like growth factor 1
IHC Immunohistochemistry
IL-1 Interleukin-1
IPMN Intraductal papillary mucinous neoplasm
LCMV Lymphocytic choriomeningitis virus
MCN Mucinous cystic neoplasm
MDSC Myeloid-derived suppressor cell
MMP Matrix metalloproteinase
MMR Mismatch repair
NF-κB Nuclear factor-κB
NSG NOD scid gamma
PanIN Pancreatic intraepithelial neoplasm
PARP Poly ADP-ribose polymerase
PDAC Pancreatic ductal adenocarcinoma
PDGF Platelet-derived growth factor
PDX Patient-derived xenograft
PSC Pancreatic stellate cell
2
RFP Red fluorescent protein
rtTA Reverse tetracycline-controlled transactivator
STAT3 Signal transducer and activator of transcription 3
TCA Tricyclic acid
TCR T cell receptor
TGFα Transforming growth factor-α
TIMP Tissue inhibitor of metalloproteinases
TNFα Tumor necrosis factor-α
TRE Tetracycline response element
TSLP Thymic stromal lymphopoietin
VEGF Vascular endothelial growth factor
I. INTRODUCTION
Background
Pancreatic ductal adenocarcinoma (PDAC; used interchangeably with pancreatic cancer
hereafter), the predominant form of pancreatic malignancy, is currently the fourth leading cause
of all cancer-related deaths in developed countries and is projected to become second only to
lung cancer by year 2024.(1) In 2015 worldwide, 367,000 patients were newly diagnosed with
pancreatic cancer, of whom 359,000 patients died due to pancreatic cancer-related causes within
the same year.(2) Although surgical resection is currently the only curative treatment for
pancreatic cancer, fewer than 20% of patients have resectable disease by the time their diagnosis
is made. The overall survival rate at 5 years is less than 7%, with most of the survivors at 5 years
belonging to the group of 10-20% of patients who undergo surgical resection of their tumors.(3)
Even for those patients undergoing surgery, 80% of them eventually relapse and die from
pancreatic cancer.
The exceptionally poor prognosis of pancreatic cancer can be attributed to several
factors.(2) First is its late diagnosis due to poor early detection, which is delayed by the absence
of clear or disease-specific symptoms and the lack of reliable biomarkers for effective screening.
Secondly, pancreatic cancer takes an aggressive course, with perineural and vascular invasions
3
and early distant metastases precluding a potentially curative surgical resection. Thirdly,
pancreatic cancer displays remarkable resistance to conventional modalities of cancer therapy,
including chemotherapy, radiotherapy as well as more recently developed molecularly targeted
therapies including immunotherapy. Finally, pancreatic cancer harbors complex tumor biology
with both intertumoral and intratumoral genetic heterogeneity, resulting in variable treatment
responses from patient to patient thus rendering a generalized approach to therapy difficult. A
comprehensive, mechanistic understanding of the pathophysiology underlying pancreatic cancer
is fundamental to overcoming these barriers.
Cell of Origin
The normal pancreas consists of two distinct functional components: endocrine and
exocrine. The endocrine component consists of glucagon-producing alpha cells and insulin-
producing beta cells that are anatomically organized into islets, and can give rise to a relatively
rarer form of pancreatic malignancies termed pancreatic neuroendocrine tumors, which have
been found to harbor mutational signatures clearly distinct from those of PDAC. These
signatures include inactivation of genes MEN1, ATRX and DAXX, derangements in the mTOR
signaling pathway, recurrent YY1 Thr372Arg missense mutations, and biallelic MUTYH
inactivating mutations.(4)
The exocrine component of the pancreas consists of digestive enzyme-secreting acinar
cells and bicarbonate-secreting ductal cells. Historically, ductal cells were thought to be the
unique source of PDAC, given their co-expression of epithelial markers, such as CK19. Recent
studies using genetically engineered mouse models of PDAC have shown that in fact both ductal
and acinar cells can give rise to PDAC precursor lesions by oncogenic KRAS activation.(4)
Furthermore, transient acinar-to-ductal metaplasia was observed in mouse models, with
4
reversible phenotypic and molecular changes that persisted in the presence of chronic
inflammation or oncogenic KRAS activation. Although there is also evidence for this
phenomenon in resected human PDAC surgical specimens, it has been argued that the
metaplastic lesions may be intraductal spread of pre-existing PDAC and/or its precursor lesions.
Genetic Landscape of Pancreatic Cancer
The genetic landscape of PDAC is characterized predominantly by mutations in four
major driver genes, listed in the order of decreasing frequency: KRAS, CDKN2A, SMAD4, and
TP53. Frequent alterations in these genes were first identified by candidate gene sequencing and
have since been corroborated repeatedly by multiple large exome and genomic sequencing
studies of PDAC.(5) Activating mutations of oncogene KRAS are seen in more than 90% of
PDACs, and inactivating mutations of tumor suppressor genes, CDKN2A, SMAD4 and TP53 in
50-80% of PDACs.(2) An additional 32 recurrent ‘passenger’ mutations – defined as those co-
occurring with driver mutations without conferring additional growth advantage – were also
identified, including but not limited to ARID1A, RNF43, TGFBR1, TGFBR2, MLL3, MKK4,
KDM6A, PREX2, RB1 and CCND1, at lower frequencies in approximately 10% of PDAC
tumors, highlighting the significance of tumoral heterogeneity (Table 1).(2, 4) It will be
important to fully characterize the functional significance of these passenger gene mutations as
they represent genetic differences among PDACs that may be exploited clinically.
Precursor Lesions
At least three histologically distinct precursor lesions of PDAC have been described so
far, consisting of pancreatic intraepithelial neoplasm (PanIN), and two types of mucinous cystic
lesions including intraductal papillary mucinous neoplasm (IPMN) and mucinous cystic
neoplasm (MCN). These precursor lesions are further characterized histologically and graded
5
according to their degree of dysplasia as lesions of low-grade versus high-grade dysplasia
(Figure 1).
Targeted sequencing of PanIN lesions along with their matched corresponding PDAC
surgical resection specimens demonstrated that the same four driver genes are mutated in PanIN
at very high frequencies as observed in PDAC. Comprehensive exome and whole genomic
sequencing studies also confirmed these findings, establishing PanIN as the canonical precursor
lesion of PDAC.(5) Similarly, shared mutations were also seen with mucinous cysts and their
matched corresponding PDACs. Targeted sequencing of IPMNs identified shared mutations in
genes GNAS and KRAS, and exome sequencing of IPMNs and MCNs identified shared
mutations in RNF43, indicating that cystic neoplasms represent additional precursor lesions of
PDAC that employ different progression pathways.(4)
Remarkably, mutational analysis comparing PanINs of different grades revealed a
positive correlation between the PanIN grade and the frequencies at which driver gene mutations
are found.(5) Furthermore, it revealed a sequential pattern in which mutations found to
accumulate in a predictive order following the PanIN grade. High-sensitivity methods to detect
KRAS mutations showed their involvement in more than 99% of all PanIN-1 lesions, suggesting
that oncogenic transformation of KRAS is most likely the initiating step in the development of
pancreatic cancer.(6) While KRAS mutations are found across all grades of PanINs and invasive
PDACs, the proportion of cells harboring the mutation increases with higher PanIN grade,
indicating a clonal expansion of cells carrying the mutation.(6) In addition to oncogenic KRAS,
inactivating mutations in CDKN2A can be seen in PanIN-2 and again at a higher frequency in
PanIN-3.(5) Similarly, SMAD4 and TP53 mutations are additionally found in PanIN-3 and in
6
invasive PDACs, with both SMAD4 and TP53 mutations occurring at higher frequencies in
invasive PDACs.
These findings may be explained by a linear progression model of pancreatic cancer
development, in which mutations are acquired in a gradual, step-wise pattern. By sequencing
primary PDACs and their matched metastatic tumors, it was estimated that the linear progression
from a nascent pancreatic cell acquiring an initiating driver gene mutation to the ultimate
development of invasive PDAC would take 10 or more years.(7) This notion is consistent with
the observation that nearly 33% of pancreata seen in autopsy series contain PanINs, suggesting
that PanINs are quite common and generally do not progress to an invasive cancer.(6) In
contrast, an alternative model termed chromothripsis proposes a punctuated evolution of
pancreatic cancer, in which catastrophic genomic events involving structural alterations cause
simultaneous inactivation of multiple driver genes in a single cell cycle. In support of this model,
whole genome sequencing of primary tumors demonstrated two-thirds of PDACs having
complex structural variations that, in a subset of cases, simultaneously inactivated multiple driver
genes.(8) In the same study, many tumors did not harbor the predicted sequence of mutations,
suggesting that these mutations may be acquired in a stochastic fashion consistent with a
chromothripsis model. Still, a third model that combines both linear progression and punctuated
evolution is entirely plausible. Distinguishing among these mechanistically distinct yet mutually
non-exclusive models has clinical importance, since under a linear progression model which
predicts a slow and gradual progression of disease, clinical efforts are best geared toward
improving methods for screening and early detection of pancreatic cancer, whereas under a
punctuated evolution model, an emphasis on enhancing systemic therapy is more appropriate.
7
Mutational Processes
To fully understand the pathophysiology of PDAC, it is essential to delineate the
mutational processes that are operative in the development of PDAC. Framing pancreatic cancer
in familiar evolutionary terms can facilitate a mechanistic understanding of how mutations arise
in the first place. In Darwinian evolution, mutations occur purely stochastically in dividing cells
at an expected somatic mutation rate of three single nucleotide variants per cell division.(6) In
the case of the pancreas which does not comprise of highly proliferative tissues, the probability
of a pancreatic cell acquiring an initiating driver gene mutation by random chance alone is
exceedingly low, and can be expected to largely depend on the total number of cell divisions
performed over the lifetime of the dividing cell. Not surprisingly, statistical analysis of various
types of human cancers, including PDAC, revealed a strong correlation between lifetime cancer
risk and the number of cell divisions performed by adult stem cells of a given organ.(9) This
finding lends support to the well-established finding that patient age is a major risk factor for the
development of PDAC. Indeed, most pancreatic cancer patients are diagnosed at beyond age 50,
with peak incidence occurring in the seventh and eighth decades of life.(2) However, the relative
contribution of intrinsic factors (e.g. stochastic mistakes taking place during DNA replication)
versus extrinsic factors (e.g. patient exposure to carcinogens or radiation) to lifetime risk remains
a point of contentious debate.
By whole-genome and RNA sequencing of resected PDAC surgical specimens, Connor
et. al identified four distinct mutational processes acting on the PDAC genome.(4, 10) Those
related to increasing age and number of cell divisions were the most prevalent, accounting for
approximately 70% of all mutational signatures observed. To lesser degrees, mismatch repair
(MMR) defects accounted for 2%, homologous recombination (HR) defects accounted for 11%,
8
and a process of unknown etiology termed “Signature 8” accounted for 15% of the mutational
signatures. Tumors with MMR and HR defects characteristically showed biallelic inactivation of
genes essential for the respective DNA repair processes, including MSH2, BRCA1, BRCA2 and
PALB2. Also, one allele was often lost in the germline, which explains the involvement of the
same genes in familial pancreatic cancers. Of note, tumors with MMR defects, owing to their
microsatellite instability, exhibited higher burdens of somatic mutations and increased
transcription of antitumor immune markers as determined by RNA sequencing, which may
translate to a greater responsiveness to immunotherapy.
Tumoral Heterogeneity
The complex genetic landscape of PDAC is complicated by significant tumoral
heterogeneity, which can be further categorized into intratumoral and intertumoral heterogeneity.
Intratumoral heterogeneity, which describes genetic heterogeneity that exists among cells
of a single tumor, is a well-recognized prognostic factor and an important cause of therapeutic
resistance in pancreatic cancer. The concept of intratumoral heterogeneity first became apparent
in lineage tracing studies of primary PDACs and matched metastatic tumors, which determined
that metastatic tumors arise from distinct subclonal outgrowths from the primary lesion, all of
which likely diverged from a single parental clone.(7) Intratumoral heterogeneity in a patient can
manifest in three forms: [1] subclonal heterogeneity within a primary tumor, where a founder
clone gives rise to various subclones by acquiring additional mutations, [2] subclonal
heterogeneity within a metastasis, where a metastasis-initiating cell gives rise to its descendant
subclones in a similar fashion, and [3] subclonal heterogeneity of metastasis-initiating cells
within a primary tumor, where metastasis-initiating cells share common ancestors but possess
distinct mutations that confer varying degrees of metastatic potential.(6)
9
Intertumoral heterogeneity describes genetic heterogeneity that exists among tumors of
same histological type occurring in different patients, and it has been well-described in
pancreatic cancer. To characterize the intertumoral differences systematically, several
classification systems have been proposed based on genomic, transcriptomic, and
immunohistochemical analyses.
Molecular Subtyping of Pancreatic Cancer
Waddell et al. classifies PDAC into four major subtypes based on patterns of structural
variation identified from their genomic analysis.(11) In their study, 20% of tumors had ‘stable’
genomes with fewer than 50 structural variants, 36% of tumors had ‘scattered’ structural events
with 50-200 variants, 14% of tumors had ‘unstable’ genomes with more than 200 structural
variants suggestive of defects in DNA maintenance, and lastly, 30% of tumors had a ‘locally
rearranged’ pattern with fewer than 50 structural variants localized to 1-3 chromosomes which
typically result from amplifications that encompass oncogenes or genomic catastrophes such as
in the case of chromothripsis. Interestingly, the ‘unstable’ subtype was predictive of platinum
and poly (ADP-ribose) polymerase (PARP) inhibitor responsiveness.
Transcriptomic studies of PDAC have also identified different molecular subtypes of
PDAC with prognostic and therapeutic implications, resulting in a number of classification
systems that differ based on the input material used and assumptions made for each study. Using
microarray expression analysis of microdissected epithelium, Collison et al. classifies PDAC into
three subtypes termed ‘classical’, ‘quasimesenchymal’ and ‘exocrine-like’.(12) Notably, the
classical subtype was predictive of therapeutic response to erlotinib, while the
quasimesenchymal subtype was negatively prognostic and predictive of therapeutic response to
gemcitabine. In a similar study, Bailey et al. analyzed transcriptomic data from bulk tissue
10
containing the tumor microenvironment, and identified an additional ‘immunogenic’ subgroup
based on presence of stromal immune cell populations.(13) Still, Moffitt et al. proposed a new
classification system by excluding transcripts from presumed normal pancreas from their
analysis, and identified two tumoral subtypes – ‘classical’ versus ‘basal-like’ – as well as two
stromal subtypes – ‘normal’ versus ‘activated’.(14) Tumors corresponding to ‘basal-like’
subtype and ‘activated’ stromal subtype were independently and additively negatively
prognostic. The basal type was also more responsive to chemotherapy on retrospective analysis.
Although large-scale genomic and transcriptomic analyses have greatly elucidated the
intertumoral heterogeneity of PDAC defining its molecular subtypes and established a
foundation for developing precision medicine, applying this knowledge clinically has been
limited by the common lack of access to complex tumor tissue biobanking and sequencing
platforms for most clinicians. To this end, Noll et al. asked whether immunohistochemical (IHC)
analysis, which is a far more accessible and technically feasible form of testing for clinicians at
large, could be used to subtype pancreatic tumors by protein expression, and determined two
IHC markers – HNF1A and KRT81 – for the differentiation of Collison subtypes.(15)
Specifically, HNF1A-positive tumors correlated to the exocrine-like subtype, KRT81-positive
tumors to quasimesenchymal subtype, and IHC-negative tumors to classical subtype. In addition,
their study identified CYP3A expression as a novel mechanism of drug resistance, found at
higher levels in exocrine-like tumors but inducible in all subtypes.
In 2009, Farrell et al. reported the predictive value of an IHC-based assay for guiding
precision medicine treatment of pancreatic cancer. In a phase III adjuvant therapy trial of 538
patients with early pancreatic cancer, the expression of human equilibrative nucleoside
transporter (hENT1) – a key mediator of cellular uptake of gemcitabine – measured by IHC
11
analysis of tumor microarrays was associated with increased overall survival and disease-free
survival in patients who received gemcitabine, but not in those who received 5-FU,
demonstrating hENT1 as a predictive biomarker for gemcitabine efficacy in patients with early
pancreatic cancer.(16)
Deranged Signaling Pathways / Molecular Aberrations
The full mutational landscape of pancreatic cancer is highly complex and diverse.
PDACs contain an average of 63 genetic alterations, the majority of which consists of infrequent
mutations found in fewer than 10% of PDACs.(2, 17) Nonetheless, many cases of these low-
frequency targets appear to be alternative perturbations of the same core signaling pathways that
are commonly deranged across all PDAC subtypes. By interrogating the exome of 24 PDACs,
Jones et al. determined 12 core signaling pathways consistent with the hallmarks of cancer
previously described by Hanahan and Weinberg, although the specific genes and the number of
genes altered in each pathway differed from patient to patient.(17, 18) Included among the
pathways were those affected by well-known driver genes, such as TP53 in DNA damage
response and SMAD4 in TGFβ signaling. Some pathways, such as RAS-ERK signaling and
DNA damage response, were predominated by a single frequently mutated gene, while others,
such as integrin signaling, regulation of invasion, homophilic cell adhesion and GTPase-
dependent signaling, involved many different genes. Biankin et al. further enriched our
knowledge of commonly deranged pathways by next-generation exome sequencing, shedding
light on the deregulation of axon guidance (SLIT and ROBO2), DNA damage repair (ATM) and
chromatin modification (EPC1) in PDAC, which were formerly unappreciated.(19)
Aberrant autocrine and paracrine signaling cascades ultimately promote pancreatic cancer
cell proliferation, migration, invasion, and metastasis.(2) Numerous cytokines, such as
12
transforming growth factor-α (TGFα), insulin-like growth factor 1 (IGF1), fibroblast growth
factors (FGFs) and hepatocyte growth factor (HGF), and their respective tyrosine kinase
receptors, lead to pathologic activation of multiple pathways that confer pancreatic cancer cell
mitogenic self-sufficiency. These signaling cascades also act to promote cancer cell migration
and invasion of both local and distant sites, leading to metastasis. Pancreatic cancer cell
proliferation is further enhanced by pathologic activation of anti-apoptotic and pro-survival
pathways, such as signal transducer and activator of transcription 3 (STAT3), nuclear factor-κB
(NF-κB) and AKT. Reactivation of genes involved in early development, such as WNT, SHH
and NOTCH, can also be seen in a subset of PDAC.
Pathway derangements in PDAC are numerous, and deconstructing their downstream
effects is further complicated by significant crosstalk between pathways creating synergistic
outcomes.(6) p53 normally cooperates with receptor SMADs to activate TGFβ-induced
transcription by forming complexes that bind separate cis-enhancer elements on a target gene
promoter. In PDAC, oncogenic KRAS interferes with TGFβ signaling by degrading SMAD4 and
inhibiting p53 by blockade of its amino-terminal phosphorylation. Furthermore, oncogenic
KRAS and mutant p53 form pathologic complexes that in turn inhibit p63, which normally acts
to oppose TGFβ-dependent cell migration, invasion and metastasis. Collectively, these findings
indicate that deranged pathways in pancreatic cancer exist not as independent processes but
rather as a complex tumorigenic network altering the systems biology of the cell.(6)
Tumor Microenvironment
A hallmark of PDAC is its abundant and dense collagenous stroma, which may account
for up to 90% of the total tumor volume. The tumor microenvironment of PDAC consists of a
highly complex assembly of diverse cell types, including pancreatic stellate cells (PSCs),
13
immune cells, endothelial cells and nerve fibers, which are influenced by the extracellular matrix
composed of matricellular proteins, fibrillar collagen, fibronectin, hyaluronic acid and a wide
range of cytokines, such as TGFβ, FGF, epidermal growth factor (EGF) receptor ligand, vascular
endothelial growth factor (VEGF) and connective tissue growth factor (CTGF).
There is now abundant evidence for the prominent role of pancreatic cancer-associated
stroma in tumor progression by actively promoting tumor growth, invasion and metastasis.
Recently, a protective effect of some of the stromal components contributing to a physical
containment of cancer cells has also been suggested. The dual function of PDAC stroma as both
a tumor promoter and a suppressor suggests that its pathogenic role may arise from a loss of
balance between epithelial cells and stroma. While normal extracellular matrix has the capacity
to restrain tumor growth through the histone demethylase JMJD1a, desmoplastic stroma consists
of aberrant matrix that is stiff with thickened collagen fibers and expresses p-MLC2 that
contributes to tumor progression.(20, 21) PSCs are major drivers of the desmoplastic reaction in
PDAC, wherein pancreatic tissue injury leads to PSC activation and trans-differentiation into α-
smooth muscle actin expressing myofibroblast-like cells secreting collagen-type I, matrix
metalloproteinases (MMPs) and tissue inhibitor of metalloproteinases (TIMPs) that remodel the
extracellular matrix. PSC activation can be triggered by various cytokines and stimuli, including
platelet-derived growth factor (PDGF), TGFβ1, FGF, EGF, tumor necrosis factor-α (TNFα),
interleukin-1 (IL-1), ethanol, endotoxins, hypoxia, pressure and oxidative stress, many of which
are produced by pancreatic cancer cells, endothelial and immune cells of the microenvironment.
Once established, PSC activation is maintained in an autocrine fashion. The resulting fibrous
stroma is a severely hypoxic, nutrient-deprived environment that promotes tumor aggressiveness
by activation of hypoxia-inducible factor-1a. In addition, activated PSCs directly promote
14
proliferation of cancer cells by secreting mitogenic factors such as stromal-derived factor-1,
PDGF, EGF, IGF-1 and FGF which activate MAPK- and AKT-signaling cascades.(22)
Another key feature of the PDAC microenvironment is its highly immunosuppressive
composition. Once the tumor is established, the tumor microenvironment is immunosuppressed
by several mechanisms, including an accumulation of regulatory T cells, M2 type tumor-
associated macrophages and myeloid-derived suppressor cells (MDSCs). Activated KRAS in
tumor cells directs the transcription of granulocyte-macrophage colony-stimulating factor (GM-
CSF), an inflammatory cytokine that promotes recruitment and trans-differentiation of myeloid
progenitor cells into MDSCs which in turn suppress the immune surveillance function of CD8+
T cells.(23) Tumor cells also stimulate the expression of IP-10 (CXCL10) in PSCs which attract
CXCR3+ regulatory T cells to the tumor milieu.(24) PSCs also secrete CXCL12 which attracts
CD8+ T cells away from the juxtatumoral stromal compartment, reducing their chance to interact
with cancer cells.(25) In addition, various cell types within the tumor microenvironment secrete
numerous cytokines that support the immunosuppressive phenotype, including IL-1b, IL-4, IL-5,
IL-6, IL-8, IL-10, IL-13, TNFα, TGFβ, FGF, PDGF, MMPs, thymic stromal lymphopoietin
(TSLP), interferon-γ (IFN-γ) and VEGF.(23) Ultimately, the PDAC microenvironment appears
to constitute a biological space of immune privilege where cancer cells are protected from
immune surveillance, as opposed to rendering T cells dysfunctional as mechanisms to bypass
mechanisms of T cell suppression can promote intratumoral infiltration of cytotoxic T cells and
uncover latent immune responses.(26, 27) Further research on the dynamic intersection of
pancreatic cancer and its tumor microenvironment is of great clinical importance as it will likely
provide answers to improving delivery of chemotherapy and developing effective
immunotherapy.
15
Metabolic Reprogramming
Successful pancreatic cancer cell survival and proliferation depends on its ability adapt to
a severely hypoxic and nutrient-deprived tumor microenvironment. Indeed, pancreatic cancer
cells are known to employ various metabolic changes through mechanisms that are mainly
driven by the expression of oncogenic KRAS and hypoxia-inducible transcription factor 1α
(HIF1α).(2) Oncogenic KRAS induces overexpression of glucose transporter 1, hexokinase 1 and
hexokinase 2, which significantly increases glucose uptake by pancreatic cancer cells. The
increased levels of glucose are funneled through aerobic glycolysis to provide substrates for ATP
production such as pyruvate as well as for the synthesis of nucleic acids, proteins, and fatty
acids. This process in PDAC is uncoupled from the tricyclic acid (TCA) cycle and electron
transport chain (ETC) via HIF1α-mediated induction of pyruvate dehydrogenase kinase 1, which
phosphorylates and inactivates pyruvate dehydrogenase, thereby limiting the conversion of
pyruvate to acetyl-CoA needed for the TCA cycle. The uncoupling of events results in increased
production of lactate, which in turn becomes an important nutrient for less hypoxic cancer cells,
and reduces the production of reactive oxygen species by ETC. Moreover, oncogenic KRAS
promotes macropinocytosis in cancer cells as a major mechanism for the uptake of extracellular
proteins to meet cellular requirements for glutamine and other amino acids. Similarly, HIF1α
activates the autophagy-lysosome system, a self-degrative process for cytoplasmic components
including organelles and macromolecules, to maintain intracellular energy supplies. In xenograft
mouse models of PDAC, pharmacologic inhibition of these processes substantially delayed
tumor growth.
16
Immune Response in Pancreatic Cancer is Unclear
Development of immunotherapies has revolutionized the treatment options for many
types of cancers, including but not limited to melanoma, renal and lung cancers. These therapies
rely on potentiating pre-existing tumor-specific T cells by blockade of immune checkpoints,
which are inhibitory pathways in place to maintain self-tolerance and modulate physiological
immune responses to minimize collateral tissue injury. The same pathways are exploited by
tumors to gain immune resistance against tumor-specific T cells. Some cancers, notably PDAC,
are refractory to immunotherapies, and it remains unclear why. The failure of numerous immune
checkpoint inhibitors to advance through clinical trials for treatment of PDAC created a
preconceived notion in the scientific community that PDACs are poorly immunogenic tumors.
However, an increasing number of studies have now shown prominent T cell infiltrates in the
vast majority of biopsies from PDAC patients and identified unique neoantigen qualities in long-
term survivors, indicating that a meaningful immune response in PDAC is achievable.(28, 29)
However, research in this area has been hampered by the lack of pre-clinical physiologic models
of PDAC that are suited to study anti-tumor immune response.
Pre-clinical Modeling of Pancreatic Cancer
“KP-C” (KrasLox-STOP-Lox-G12D
; P53Lox-STOP-Lox-R172H/+
; Pdx1-Cre) mice have been widely
used to investigate pancreatic cancer biology. Although this model has been greatly informative
regarding the genetic landscape of PDAC, it is ill-suited for the study of cancer immunology on
two levels. First, tumors develop aggressively in these mice, rapidly progressing to fatal
metastatic disease predominantly by 6 weeks of life. This creates a practical challenge in
investigating early disease when meaningful tumor-immune cell interactions may occur before
significant stromal development and/or the onset of other mechanisms of immune
17
suppression.(30) Secondly, pancreatic tumors that develop in these mice are poorly antigenic,
lacking neoantigen peptides which are critical for mounting anti-tumor T cell responses. In fact,
depletion of T cells in KP-C mice using anti-CD4 and anti-CD8 antibodies had no effect on the
progression of murine PDAC nor on the overall survival of these mice.(31) Thus, most
pancreatic cancer immunology studies have focused on murine and human PDAC cell lines,
which have their own limitations.(32) Namely, monolayer cell lines lack the structural
sophistication and functional differentiation of cells seen in vivo, and cannot recapitulate the
tumor microenvironment in mouse xenograft studies. Cell line-derived three-dimensional
spheroid cultures attempt to address this issue, but are difficult to propagate in spheroid form,
limiting longitudinal investigations. Furthermore, none of the cell-line derived models support
the growth of untransformed, non-neoplastic cells. Instead, they inevitably become monoclonal
over time by in vitro selection of the most aggressive clones, resulting in a loss of genetic
heterogeneity seen in primary tumors. Patient-derived xenograft (PDX) mouse models, which are
established by implanting a piece of surgically resected tissue from a patient under the dermis of
immunocompromised mouse hosts, are inherently more physiologic but are cost-prohibitive and
excessively time-consuming, commonly taking upwards of 6 months to generate sufficient sizes
of mouse colonies, which is outside clinically meaningful timeframes for any approach to
personalized medicine for most pancreatic cancer patients.
A recent breakthrough in translational pancreatic cancer research has been the
development of organoid models of pancreas using human and mouse pancreatic tissues for pre-
clinical modeling of PDAC. Organoids, comprising of complex clusters of multiple cell types
derived from the tissue of interest, can recapitulate the intricate spatial architecture of the
progenitor organ structure and perform functions of the organ such as secretion or contraction.
18
Since a robust method for production of self-renewing intestinal organoids was first reported in
2009, tumor organoid models have been widely adopted for multiple organ systems.(33) In 2015,
Boj et al. recently described methods for reliably generating human and mouse PDAC organoids
using surgical resection specimens as well as endoscopic ultrasound-guided fine needle biopsy
(EUS-FNB) specimens.(34) PDAC organoids derived in this manner could recapitulate the
natural history of human PDAC when orthotopically transplanted into immunocompromised
mice, forming early PanIN-like lesions that progressed to invasive pancreatic cancer with robust
stromal response. The ability to generate organoid cultures from FNB specimens is a major
advantage, since it enables investigators to capture the full spectrum of PDAC ranging from
early premalignant lesions to late metastatic cancers, as opposed to surgical resection specimens
which account for fewer than 20% of patients diagnosed with PDAC who are surgical
candidates. The organoid model is physiologic yet possesses all the desirable intrinsic properties
of an in vitro system. PDAC organoid cultures can be propagated in vitro for expansion of
starting material, which is often the limiting factor for tissue-consuming studies such as deep
sequencing, and cryopreserved indefinitely without losing genetic heterogeneity. They are highly
tractable, amenable to genetic manipulation and high-throughput assays. Moreover, in contrast to
PDX mouse models, organoid cultures can be established rapidly in sufficient quantities for
studies in just 2-4 weeks from the time point of acquiring patient tissues, permitting a
personalized approach to pancreatic cancer medicine to investigate patient-specific tumor
biology, evaluate prognosis and guide therapy in real time.
II. STATEMENT OF PURPOSE
In this study, we sought to develop murine and human organoid models of PDAC to
investigate the biology of pancreatic cancer immune response. Our aims were mainly two-fold:
19
1. Development of an immunogenic murine PDAC organoid model to study antigen-
specific anti-tumor T cell responses in both in vivo and in vitro setting.
2. Creation of a clinically annotated library of validated, patient-derived PDAC organoid
lines as tools for studying human pancreatic cancer immunology.
III. METHODS
Acquisition of human specimens
Human pancreatic cancer tissues were obtained from patients undergoing endoscopic
ultrasound-guided fine needle biopsy (EUS-FNB) or surgical resection at Yale New Haven
Hospital. Some of the surgical resection specimens were used to create patient-derived xenograft
(PDX) mouse models, which subsequently became available as a secondary source of patient-
derived tissues for generation of organoids. Tissues were determined to be tumoral or normal by
evaluation of on-site clinical pathologist. Written informed consent was obtained from all
patients prior to tissue acquisition. This study was reviewed and approved by the Institutional
Review Board of Yale University. All EUS-FNB specimens were provided by James Farrell who
also performed the biopsies. All surgical resection specimens were histologically evaluated and
provided by Marie Robert. PDX mouse models of PDAC were previously established by Ryan
Sowell in Kaech laboratory.
Isolation and culture of murine pancreatic organoids
Murine pancreatic organoids were generated using normal or pre-neoplastic pancreatic
tissues from C57BL/6 mouse and KP-NINJA (KrasLox-STOP-Lox-G12D
; P53flox/flox
; inversion induced
joined neoantigen) mouse, respectively. Detailed procedures for isolation and propagation of
murine pancreatic organoids were adapted from Boj et al., 2015 and Huch et al., 2016. Briefly,
mouse pancreas was dissected and minced into sub-millimeter pieces before enzymatic digestion
20
with collagenase XI (0.125 mg/mL, Sigma-Aldrich), dispase II (0.125 mg/mL, Thermo
Scientific) and DNase I (0.1 mg/mL, Sigma-Aldrich) in advanced DMEM/F12 medium (Life
Technologies) supplemented with FBS (2.5%), Glutamax (1X, Thermo Scientific) and
Antibiotic-Antimycotic (1X, Thermo Scientific) for 1-3 hours at 37ºC in a tissue dissociator until
visual confirmation of pancreatic ducts which were manually picked for ductal enrichment under
a dissecting microscope. Harvested ductal fragments were embedded in growth factor reduced
Matrigel (Corning) and cultured in complete murine organoid growth medium, consisting of
advanced DMEM/F12 supplemented with Glutamax (1X), HEPES (10 mM, Life Technologies)
and Antibiotic-Antimycotic (1X), Rspo1-conditioned medium (10% v/v), human noggin (0.1
µg/mL, Peprotech), B27 supplement minus vitamin A (1X, Thermo Scientific), N-acetyl cysteine
(1.25 mM, Sigma-Aldrich), nicotinamide (10 mM, Sigma-Aldrich), human gastrin I (10 nM,
Sigma-Aldrich), mouse EGF (50 ng/mL, Thermo Scientific), human FGF-10 (100 ng/mL,
Peprotech) and A83-01 (500 nM, Tocris Bioscience). Y-27632 (10.5 µM, Tocris Bioscience)
was added for initial organoid cultures following isolation from primary tissue, single cell
dissociation, or thawing from cryopreservation. Murine organoid models were characterized by
in vivo transfer for tumor formation in C57BL/6 mouse and immunohistochemical analysis of
resulting tumors. KP-NINJA mouse model was previously established by Nikhil Joshi. All
procedures outlined above were performed by the author.
Isolation and culture of human PDAC organoids
Human PDAC organoids were generated using patient-derived tissues from EUS-FNB,
surgical resection, or pre-established PDX mouse models. Detailed procedures for isolation and
propagation of human PDAC organoids were adapted from Boj et al., 2015 and Huch et al.,
2016. Briefly, tissues were minced into sub-millimeter pieces before enzymatic digestion with
21
collagenase II (5 mg/mL, Thermo Scientific), dispase II (0.125 mg/mL) and DNase I (0.1
mg/mL) in advanced DMEM/F12 medium supplemented with FBS (2.5%), Glutamax (1X) and
Antibiotic-Antimycotic (1X) for 1-3 hours at 37ºC in a tissue dissociator until tissues become
submacroscopic. Cells are embedded in growth factor reduced Matrigel and cultured in complete
human organoid growth medium, consisting of advanced DMEM/F12 supplemented with
Glutamax (1X), HEPES (10 mM) and Antibiotic-Antimycotic (1X), Wnt3a-conditioned medium
(50% v/v), Rspo1-conditioned medium (10% v/v), human noggin (0.1 µg/mL), N2 supplement
(1X, Thermo Scientific), B27 supplement minus vitamin A (1X), N-acetyl cysteine (1.25 mM),
nicotinamide (10 mM), human gastrin I (10 nM), human EGF (50 ng/mL, Peprotech), human
FGF-10 (100 ng/mL, Peprotech) and A83-01 (500 nM). Y-27632 (10.5 µM) was additionally
added for initial organoid cultures following isolation from primary tissue, single cell
dissociation, or thawing from cryopreservation. Human organoid models were characterized by
Sanger sequencing of KRAS and P53, in vivo transfer for tumor formation in immunodeficient
NOD scid gamma (NSG) mouse (The Jackson Laboratory) and immunohistochemical analysis of
resulting tumors. FNB-derived organoid lines were established by the author. Surgical resection-
derived and PDX mouse-derived organoids were established by collaborative effort of Prashanth
Gokare and the author. Characterization of patient-derived organoids was performed by
collaborative effort of Prashanth Gokare and the author.
Isolation of primary murine PDAC cell lines
Primary murine PDAC cell lines were isolated from pancreatic tumors harvested from
KP-C (KrasLox-STOP-Lox-G12D
; P53flox/flox
; Pdx1-Cre) mouse. Resected tumors were minced into
sub-millimeter pieces before enzymatic digestion with trypsin-EDTA (0.25%, Life
Technologies) and collagenase IV (1 mg/mL, Worthington Biochemical) in HBSS buffer (1X,
22
Life Technologies) for 30 min at 37ºC in a tissue dissociator, after which the digestion reaction
was quenched using cold FBS. Cells were passed through a 40 µm filter to prepare single cell
suspension and washed twice prior to cell culture in complete DMEM. All procedures outlined
above were performed by Brittany Fitzgerald.
Genetic manipulation of murine pancreatic organoids
Murine pre-neoplastic pancreatic organoids isolated from KP-NINJA mouse model were
in vitro transformed into neoplastic organoids using LV-rtTA-Cre-iRFP670. Detailed procedures
for genetic manipulation of organoids were adapted from Huch et al., 2016. Briefly, a single cell
suspension of organoids was prepared by pooling 3 confluent wells of a 24-well plate, removal
of Matrigel, and digestion of organoids in TrypLE Express (1X, Life Technologies) and DNAse I
(0.1 mg/mL) for 5 min at 37ºC with vigorous pipetting every 2 min. After washing, cells were
resuspended with concentrated lentivirus and spinoculated at 600 G for 1 hour at 32ºC, followed
by incubation for 6 hours at 37ºC. iRFP670 labeling of infected organoid fragments could be
visualized by fluorescence microscopy 2-3 days after infection. After expansion, organoids were
analyzed by flow cytometry for expression of iRFP670 and sorted for the brightest 10% of cells
expressing iRFP670. For the expression of programmed, GFP-tagged neoantigens, a subset of
transformed organoids was also infected with Ad-FLPo by spinoculation followed by incubation
as described above. Ad-Cre was used as a negative control. Alternatively, a subset of organoids
was treated with doxycycline and tamoxifen in vitro to achieve the same effect. After expansion,
organoids were analyzed by flow cytometry for expression of GFP and sorted for the brightest
10% of cells expressing GFP. Lentiviral and adenoviral transformations of organoids were
performed by the author. Creation of immunogenic organoid lines by treatment with doxycycline
23
and tamoxifen was performed by Gena Foster. Flow cytometry and cell sorting of transformed
organoids were performed by collaborative effort of Gena Foster and the author.
In vivo mouse assays
Murine and human pancreatic organoids were characterized by subcutaneous injection of
organoids for in vivo tumor formation in C57BL/6 mouse or NSG mouse, respectively. To
standardize injections, organoids were first dissociated into single cells and seeded at
concentration of 2.5 x104
cells per well in 24-well plate format. Organoids were then expanded
to 80-90% confluency in a period of 5-10 days depending on the organoid line. For each mouse
injection, organoids were pooled from 6 confluent wells for a total of approximately 5 x 105
cells, broken down into organoid fragments by vigorous pipetting using 200 µL pipette tips, and
finally resuspended in 50 µL of Matrigel diluted 1:1 with cold PBS. Mice were anesthetized
using isoflurane for injections and subsequently monitored for subcutaneous growth of tumors by
caliper measurement every 2 days. Mice were euthanized promptly when tumors reached 1 cm in
size or whenever a humane concern developed. Resulting tumors were harvested and analyzed
by immunohistochemistry. All procedures outlined above were performed by the author.
To test the effects of programmed neoantigens on tumor growth in vivo, neoantigen-
negative, neoantigen-positive and weakly neoantigen-positive murine PDAC organoids were
injected subcutaneously into C57BL/6 mice as described above. For creation of a weakly
neoantigen-positive organoid line with only 10% of its cells expressing neoantigens, neoantigen-
positive organoids were diluted 1:9 with neoantigen-negative organoids. A cohort of mice also
received retroorbital injections of luciferase-positive P14 splenocytes 24 hours prior to receiving
organoid injection for co-transfer of antigen-specific T cells. Splenocytes were harvested from
luciferase+ P14 mouse strain by homogenizing dissected spleen through a 70 µm strainer and
24
passing through a 27-gauge needle for single cell dissociation. Following centrifugation, RBC
lysis was performed by incubation of cells in ACK Lysing Buffer (1X, Thermo Scientific) for 3-
5 min at room temperature. Cells were washed twice using RPMI medium prior to flow
cytometry analysis for confirmation of tetramer-positive P14 CD8+ T cell population. For each
mouse injection, 1 x 106
splenocytes were finally resuspended in PBS. Mice were anesthetized
using isoflurane for each injection and subsequently monitored for subcutaneous growth of
tumors by caliper measurement every 2 days. Mice receiving co-transfer of organoids and
splenocytes were additionally monitored by IVIS Spectrum In Vivo Imaging System
(PerkinElmer) for luciferase detection 24 hours after organoid injection and every 3 days
thereafter. Mice were euthanized when tumors reached 1 cm in size or whenever a humane
concern developed. Resulting tumors were harvested and analyzed by immunohistochemistry.
Creation of immunogenic murine organoid lines and immunohistochemical analyses of resulting
tumors were performed by collaborative effort of Gena Foster and the author. Isolation, flow
cytometry and co-transfer of splenocytes were performed by collaborative effort of Brittany
Fitzgerald and the author.
Immunohistochemical analysis of tumors
Primary and organoid-derived pancreatic tumors were analyzed by
immunohistochemistry. Tissues were fixed in 10% formalin and embedded in paraffin. Sections
were subject to H&E, RFP (600-401-379, Rockland) 1:1000, E-Cadherin (610182, BD
Bioscience) 1:500, CK19 (Troma III, developed by Rolf Kemler, Max-Planck Institute of
Immunobiology, Freiberg, Germany, and obtained from the Hybridoma Bank at the University
of Iowa, Iowa City, Iowa, USA) 1:1000, Sox9 (AB5535, EMD Millipore) 1:1000, Muc5AC
(ab212636, Abcam) 1:400, Phospho-Erk (4370S, Cell Signaling Technology) 1:400, and
25
Phospho-Mek (2338S, Cell Signaling Technology) 1:50. Immunohistochemical staining and
imaging of tumor sections were performed by collaborative effort of Gena Foster and the author.
Sanger sequencing of organoids
Patient-derived organoids were sequenced for characteristic mutations in genes KRAS
and P53 by Sanger sequencing as part of validation pipeline. Genomic DNA was prepared from
organoids using DNeasy Blood & Tissue Kit (QIAGEN) and quantified using Nanodrop
spectrophotometer. Regions of gene that are most frequently mutated were PCR amplified and
sequenced using the same set of primers. Mutations at codons 12 and 13 of KRAS were
determined by using sense primer: 5’AAAGGTACTGGTGGAGTATTTGATAG and antisense
primer: 5’ACAAGATTTACCTCTATTGTTGGATC. Mutations at codon 61 of KRAS were
determined by sense primer: 5’GGAAGCAAGTAGTAATTGATGGAGA and antisense primer:
5’GCATGGCATTAGCAAAGACTCA. Mutations in exon 5 of P53 were determined using
sense primer: CAAGCAGTCACAGCACATGA and antisense primer:
AACCAGCCCTGTCGTCTCT. Mutations in exon 6 of P53 were determined using sense
primer: CAGGCCTCTGATTCCTCACT and antisense primer:
AGACCTCAGGCGGCTCATAG. Mutations in exon 7 of P53 were determined using sense
primer: ATCTCCTAGGTTGGCTCTGA and antisense primer:
TGGCAAGTGGCTCCTGACCT. Mutations in exon 8 of P53 were determined using sense
primer: CTCTTTTCCTATCCTGAGTA and antisense primer: CTGCTTGCTTACCCTGCTTA.
PCR products were purified using QIAquick PCR Purification Kit (QIAGEN) and analyzed by
gel electrophoresis for correct band size prior to sequencing. All procedures outlined above were
performed by collaborative effort of Prashanth Gokare and the author.
26
Development of organoid-T cell co-culture model systems
P14 CD8+ T cells were pre-activated and expanded out from harvested P14 mouse
splenocytes by incubation of cells with GP33 peptide (0.1 nM, Anaspec) for 1 hour followed by
incubation with human IL-2 (10 ng/mL, Peprotech) for 72 hours at 37ºC in complete RPMI
medium supplemented with FBS (10%), HEPES (1X), non-essential amino acids (1X, Life
Technologies), sodium pyruvate (1X, Life Technologies), 2-mercaptoethanol (55 µM, Sigma-
Aldrich), penicillin-streptomycin (1X, Life Technologies) and Glutamax (1X), followed by
cytometry confirmation and cell sorting of tetramer-positive P14 CD8+ T cells. Prior to co-
culture, murine PDAC organoids were labeled with Calcein blue, AM (1X, Anaspec) while P14
CD8+ T cells were doubly labeled with Calcein blue, AM as well as Calcein green, AM (1X,
Invitrogen). Neoantigen-positive and neoantigen-negative organoids were seeded in
10/20/30/40/50 uL volumes of Matrigel in 24-well plate format and cultured to 30% confluency.
Prepared T cells were resuspended in complete RPMI medium at a concentration of 1 x 105
cells
per 500 uL per well and were added carefully on top of Matrigel plugs after the removal of
organoid growth medium. Co-cultures were subsequently monitored by live chambered
fluorescence imaging on EVOS Cell Imaging System (Invitrogen) for 24 hours. All procedures
outlined above were performed by collaborative effort of Prashanth Gokare and the author.
IV. RESULTS
KP-NINJA mouse model provides substrate for creation of immunogenic murine organoid
models of PDAC
To overcome the paucity of neoantigen peptides on pancreatic tumors that develop in
standard KP-C mouse model and create an immunogenic murine PDAC organoid model, we
generated pancreatic organoids from “KP-NINJA” (KrasLox-STOP-Lox-G12D
; P53flox/flox
; inversion
27
induced joined neoantigen) mouse model that has been genetically engineered to express
glycoproteins GP33-41 and GP61-80 derived from lymphocytic choriomeningitis virus (LCMV)
as CD8+ and CD4+ T cell neoantigens, respectively (Figure 2A). The neoantigens are tagged to
the C-terminus of green fluorescent protein (GFP) functioning as a reporter for neoantigen
expression. In order to ensure tight regulation of its expression, multi-layered genetic and drug-
inducible mechanisms were engineered. This is critical as leaky expression of neoantigens during
early developmental phase of mouse immune system can result in immune tolerance and loss of
immunogenicity. The neoantigen cassette is inverted and flanked by non-compatible flippase
recognition target (FRT) sites, requiring the action of flippase (FLP) recombinase to be properly
expressed. The expression of FLP recombinase is regulated by a tetracycline response element
(TRE), which requires reverse tetracycline-controlled transactivator (rtTA) – which can be
introduced by any tissue specific promoter – plus doxycycline to be transcribed. The entire TRE-
FLP recombinase cassette is floxed, requiring Cre recombinase mediated inversion to become
poised for transcription. Introduction of Cre recombinase will also recombine KRAS and P53
resulting in the activation of oncogenic KRAS and deletion of P53 to drive tumorigenesis.
Finally, FLP recombinase is fused to a mutated ligand binding domain of the human estrogen
receptor (ER), requiring tamoxiphen to become stabilized and effective in the nucleus. As a
result, KP-NINJA mouse model enables genetically and pharmacologically inducible expression
of known neoantigens with precise temporal and spatial control.
To generate normal pancreatic organoids from KP-NINJA mouse model, we adapted
methods previously described by Boj et al.(34) Briefly, mouse pancreas dissection is followed by
mechanical and enzymatic digestion to release ductal fragments, which are manually picked
under a dissecting microscope for ductal enrichment (Figure 2B). The enriched ductal fragments
28
are then washed and seeded in basement membrane matrix (Matrigel) in 24-well plate format.
Liquid medium containing essential components for pancreatic organoid culture is then added on
top of congealed Matrigel plugs. After 5-10 days of tissue culture, budding of ductal fragments
into spherical organoids can be observed.
In vitro transformed murine pancreatic organoids form tumors that are histologically
similar to early lesions found in human PDAC
For in vitro transformation of KP-NINJA mouse-derived normal pancreatic organoids, a
lentiviral construct encoding rtTA-Cre-iRFP670 was used for Cre recombinase-mediated
activation of KRAS oncogene and deletion of P53. iRFP670 was included in the lentiviral
construct as a fluorescent reporter for the expression of Cre recombinase, thereby labeling any
transformed cell. Following lentiviral transformation, organoids were analyzed by fluorescence
imaging and flow cytometry for expression of iRFP670 and sorted for the brightest 10% of cells
expressing red fluorescent protein (RFP) (Figure 3A-B). Expression of GFP was included in the
flow cytometry analysis to examine the possibility of undesirable leakiness of neoantigen
expression, which did not occur in organoids.
When lentivirus-transformed organoids versus untransformed normal pancreatic
organoids were injected subcutaneously into the opposite flanks of same mouse, only the
transformed organoids formed a tumor (Figure 3C). On histology, these tumors had numerous
infiltrating well-differentiated ductal structures with epithelial lesions consisting of tall columnar
cells with mucinous cytoplasm that are reminiscent of early lesions seen in human PDAC, as
well as a robust stromal response with extensive fibrosis (Figure 3D). The ductal structures
stained positively for RFP confirming transformed organoids as their cell of origin (Figure 4C).
The stromal compartment did not stain for RFP, indicating that the stromal response is host-
29
derived. The ductal structures also stained positively for markers of epithelial and
pancreaticobiliary origin, including E-cadherin, CK19 and Sox9, as well as for PDAC-associated
tumor markers, such as Muc5AC and phosphorylated Erk and phosphorylated Mek which are
downstream targets of oncogene KRAS in the MAPK/ERK signaling pathway.
The histology of the organoid-derived tumors was comparable to that of primary murine
pancreatic tumors that spontaneously form in KP-C mice (Figure 3D-E). Murine PDAC cell
lines were generated from the primary pancreatic tumors of KP-C mice for comparative analysis.
Tumors derived from injection of PDAC cell lines showed markedly different histology to that of
tumors derived from injection of transformed organoids, notable for the absence of organized
ductal structures in cell line-derived tumors, consistent with highly advanced, undifferentiated
pathology as a result of a known caveat with monolayer cell lines, that is in vitro selection of
aggressive clones (Figure 3F). Cell line-derived tumors also lacked a stromal response in
contrast to organoid-derived tumors.
Serial in vivo transfer of transformed murine pancreatic organoids results in progressively
more advanced tumors
We predicted that serial in vivo transfer of transformed organoids in mice by performing
repeated rounds of organoid generation from tumors derived from organoid injections in a
sequential fashion would lead to progressively more advanced tumors. After two rounds of in
vivo transfer, the organoid-derived tumors showed increased features of high-grade dysplasia,
including enlarged, hyperchromatic nuclei with prominent nucleoli, nuclear crowding and cell
stacking (Figure 4A-B). Fluorescence imaging of organoids generated from tumors after one
round of in vivo transfer not only confirmed retention of RFP label but also showed a more
uniform labeling of organoids indicating in vivo enrichment for Cre recombinase-transformed
30
cells (Figure 4D). After third round of in vivo transfer, the tumors contained noticeably fewer
organized ductal structures and appeared poorly differentiated.
Expression of neoantigens in murine PDAC organoids elicits effective immune response in
mouse
For creation of immunogenic murine PDAC organoid lines, an adenoviral construct
encoding FLP recombinase was used to genetically induce expression of GFP-tagged
neoantigens in transformed organoids (Figure 5A). After adenoviral introduction of FLP
recombinase, the organoids were analyzed by flow cytometry to confirm GFP expression and
were sorted for the brightest 10% of cells expressing GFP (Figure 5B). An adenoviral construct
encoding Cre was used as a negative control. To test whether expression of neoantigens can
impact tumor growth, neoantigen positive versus neoantigen negative transformed organoids
were injected subcutaneously into 3 cohorts of mice, where first cohort received neoantigen
negative organoids, second cohort received neoantigen positive organoids, and third cohort
received neoantigen positive organoids as well as retroorbital injections of luciferase-positive
P14 CD8+ T cells, which have been genetically engineered to express the T cell receptor (TCR)
specific for GP33, 24 hours prior to organoid injections. In vivo imaging of mice at 24 hours
after organoid injections demonstrated accumulation of luciferase-positive P14 CD8+ T cells at
the site of organoid injection in the third cohort (Figure 5C). Mice were monitored for
subcutaneous growth of tumors for up to 30 days. None of the mice that received neoantigen-
positive organoids developed tumors, while all 5 out of 5 mice that received neoantigen-negative
organoids developed tumors, indicating immune clearance of neoantigen-expressing PDAC
organoids (Figure 5D). Interestingly, upon injection of a mixture of neoantigen-positive and
31
neoantigen-negative organoids at a ratio of 1:9, tumors were able to form but were heavily
infiltrated with immune cells on histology (Figure 6).
Expression of neoantigens in murine PDAC organoids promotes T cell infiltration in T cell-
organoid co-culture model
Next, we developed a novel three-dimensional co-culture system to study interactions of
murine T cells and PDAC organoids in vitro. Splenocytes were harvested from congenic P14
mice that have been genetically engineered to express TCRs specific for GP33, and were treated
with IL-2 and GP33 peptide for expansion and pre-activation of constituent P14 T cells.
Splenocytes were then analyzed by flow cytometry and sorted to prepare a pure population of
pre-activated P14 CD8+ T cells. Cell-permeant live-cell staining dyes were used to distinguish
cells in co-culture. PDAC organoids were labeled with green-fluorescent calcein AM dye, while
P14 CD8+ T cells were doubly labeled with green- and blue-fluorescent calcein AM dyes. P14
CD8+ T cells were then introduced into the liquid medium of either neoantigen-positive or
neoantigen-negative PDAC organoid cultures which were maintained in Matrigel plugs of
different sizes to vary the amount of liquid media-Matrigel interface. Co-cultures were
subsequently monitored under live fluorescence imaging for 24 hours. Within the first hour,
there was significant clustering of P14 CD8+ T cells at the boundaries of Matrigel plugs which
appeared to be a physical barrier to T cell entry (Figure 7A). Nonetheless, small yet increasingly
large fractions of T cells could be observed to penetrate the Matrigel plugs containing
neoantigen-positive PDAC organoids, such that by 24 hours there was a clearly noticeable
difference in the amount of T cell infiltration between neoantigen-positive versus neoantigen-
negative PDAC organoid co-cultures (Figure 7B).
32
Assembly of human PDAC organoid library
To aid investigations of human pancreatic cancer immunology, we sought to build a
clinically-annotated library of experimentally validated, patient-derived PDAC organoid lines at
Yale. We obtained patient samples primarily from EUS-FNB specimens but also a smaller
number of samples from surgical resection specimens and pre-established PDX mouse models of
PDAC (Figure 8A). From October 2017 to May 2018, we successfully generated 21 patient-
derived organoid lines from 24 FNB specimens, including one liver metastasis, all of which were
pathology confirmed as PDAC for an overall organoid isolation efficiency of 87.5% (21/24).
Established organoid cultures were validated as tumor organoids as opposed to normal pancreatic
contaminants by in vivo transfer of organoids for tumor formation in immunocompromised
(NSG) mouse, immunohistochemical analysis of organoid-derived tumors for common markers
of PDAC, and Sanger sequencing of organoids for KRAS and P53, the two most commonly
mutated genes in PDAC (Figure 8B-E). Under these criteria, 7 out of 9 patient-derived organoid
lines tested to date have been successfully validated.
Histology of tumors derived from organoid injections in mice closely matched the
histology of their corresponding patient-derived primary tissues, confirming that organoids truly
recapitulate the pathology of their source material (Figure 8C). The degree of dysplasia seen in
organoid-derived tumors also correlated with the severity of disease of corresponding patients at
the time of biopsy in three case studies (Fig. 8D). Organoids derived from a patient who had
borderline resectable disease (Bx120817) formed tumors of moderately differentiated histology
in agreement with the primary clinical pathology findings. In comparison, organoids derived
from the primary tumor of a patient who had metastatic disease in the lungs (Bx111417) formed
less differentiated tumors consisting of numerous disorganized ductal structures with
33
characteristic loss of lumens. Organoids derived from a metastatic lesion in the liver of a patient
who had liver metastasis (Bx102417) formed the most poorly differentiated tumors with little to
zero resemblance of normal ductal structures.
V. DISCUSSION
Development of effective immunomodulating therapies in pancreatic cancer remains
elusive, largely owing to the lack of effective physiologic PDAC model systems for the study of
biology of immune response against tumor. To fulfill this critical need, we have developed
transplantable, immunogenic murine organoid models of PDAC that enable investigations of
antigen-specific, anti-tumor T cell responses. By genetic engineering of inducible mutations in
KRAS and P53, we were able to recreate the earliest genetic events in pancreatic tumorigenesis
in vitro, and then follow the progression of disease in vivo after transplantations of organoids in
mouse. Tumors derived from organoids in this way are histologically similar to early lesions
found in human PDAC, demonstrating well-differentiated ductal structures infiltrating an
extensive and dense fibrous stroma. Ability to recapitulate the tumor microenvironment which is
fundamental to PDAC pathophysiology is essential for studying how stromal components impact
immune response. Furthermore, by serial in vivo transfer of transformed murine organoids, we
were able to generate transplantable organoid models that can reliably recreate discrete stages of
PDAC progression from early precursor lesions to advanced invasive cancer. Recreation of early
disease is especially critical for capturing meaningful tumor-immune cell interactions as
immunosuppression is an early event in PDAC.(30) By introducing known T cell neoantigens in
murine PDAC organoids, we were able to elicit robust and effective immune responses against
neoantigen-expressing PDAC organoids in mouse transplant studies. High level of neoantigen
expression in 100% of cells comprising PDAC organoids resulted in complete immune rejection
34
of organoid-derived tumor growth in mouse, whereas a low level of neoantigen expression by
dilution of neoantigen-positive cells with neoantigen-negative cells by a factor of 10 permitted
tumor growth albeit with increased immune infiltration, suggesting that both quality and quantity
of neoantigen affect immune response. The expression of high-quality neoantigens in sufficient
quantity before the development of an immunosuppressive tumor microenvironment may have
been key to the successful immune clearance of PDAC organoids in this set-up. To facilitate
mechanistic studies of antigen-specific T cell responses against tumor, we have developed an in
vitro three-dimensional co-culture system that recapitulates interactions of T cells and PDAC
organoids in vitro, where increased T cell infiltration of Matrigel plugs containing neoantigen-
expressing organoids was observed. This system can be readily applied to study tumor organoid
interactions with other important cell types such as tumor-associated macrophages or cancer-
associated fibroblasts. Ability to generate effective immune responses against tumor using
organoid models of PDAC challenges the widely conceived notion that PDAC is an inherently
immunologically cold disease.
Precision medicine is a newly emerging medical model that accounts for the unique
biology of each patient, allowing for the development of targeted therapeutics against specific
molecular mechanisms at play and a personalized approach to disease management based on the
individual patient’s tumor characteristics. Since the advent of next-generation sequencing, our
knowledge of biomolecular and genetic aspects of pancreatic cancer has grown exponentially in
recent years, revealing novel insights into how precision medicine may be actualized, including
an appreciation for the remarkable heterogeneity of PDAC. Integrated analyses of different
‘omics’ data sets enabled the categorization of pancreatic tumors into distinct molecular subtypes
carrying prognostic and predictive values, paving the way to individualized therapy by
35
stratifying patients based on their tumor subtype for specific therapies. However, personalized
medicine in pancreatic cancer has been difficult achieve due to the short median survival of
pancreatic cancer patients and long turnaround times of standard PDX models. Development of
patient-derived organoid models of PDAC has been revolutionary in this regard, as organoids
can be derived from patients rapidly and analyzed within clinically relevant timeframes. In some
cases, pharmacotyping of patient-derived organoids to generate drug-sensitivity profiles could be
completed in as little as 6 weeks.(35) Moreover, organoids can model the full clinical spectrum
of PDAC as they can be generated using small amounts of tissue from FNB specimens, thus
removing the barrier to sampling non-surgical patients who account for more than 80% all
pancreatic cancer patients, as opposed to standard PDX models that require surgical tissues. In
our study, we have established a clinically annotated library of 30+ patient-derived PDAC
organoid lines using FNB and surgical specimens. Our efforts to validate each patient-derived
organoid line by tumor formation in mouse, immunohistochemistry and sequencing have been
promising.
Collectively, our data demonstrate that pancreatic organoids are an ideal model for the
study of pancreatic cancer immune response. Our ongoing work includes using CRISPR/Cas9-
based lentiviral systems in PDAC organoids to test and define genes that impact anti-tumor T
cell responses with or without addition of immunomodulating agents in both in vitro co-cultures
and in vivo mouse studies. We are also using organoid models of PDAC to investigate the
pathogenic role of renalase – a recently discovered cytoprotective secreted flavoprotein that is
upregulated in chronic pancreatitis and PDAC – and evaluating its potential use as both
predictive biomarker and a therapeutic target.(36) Continuation of efforts using organoid models
of PDAC to understand the mechanistic underpinnings of immunomodulating therapies and
36
advance research in pancreatic cancer early detection and precision medicine should accelerate
improvement of patient outcomes for this deadly disease.
VI. REFERENCES
1. Rahib L, Smith BD, Aizenberg R, Rosenzweig AB, Fleshman JM, and Matrisian LM.
Projecting cancer incidence and deaths to 2030: the unexpected burden of thyroid, liver,
and pancreas cancers in the United States. Cancer Res. 2014;74(11):2913-21.
2. Kleeff J, Korc M, Apte M, La Vecchia C, Johnson CD, Biankin AV, et al. Pancreatic
cancer. Nat Rev Dis Primers. 2016;2:16022.
3. He J, Ahuja N, Makary MA, Cameron JL, Eckhauser FE, Choti MA, et al. 2564 resected
periampullary adenocarcinomas at a single institution: trends over three decades. HPB
(Oxford). 2014;16(1):83-90.
4. Oldfield LE, Connor AA, and Gallinger S. Molecular Events in the Natural History of
Pancreatic Cancer. Trends Cancer. 2017;3(5):336-46.
5. Hruban RH, Goggins M, Parsons J, and Kern SE. Progression model for pancreatic
cancer. Clin Cancer Res. 2000;6(8):2969-72.
6. Makohon-Moore A, and Iacobuzio-Donahue CA. Pancreatic cancer biology and genetics
from an evolutionary perspective. Nat Rev Cancer. 2016;16(9):553-65.
7. Yachida S, Jones S, Bozic I, Antal T, Leary R, Fu B, et al. Distant metastasis occurs late
during the genetic evolution of pancreatic cancer. Nature. 2010;467(7319):1114-7.
8. Notta F, Chan-Seng-Yue M, Lemire M, Li Y, Wilson GW, Connor AA, et al. A renewed
model of pancreatic cancer evolution based on genomic rearrangement patterns. Nature.
2016;538(7625):378-82.
9. Tomasetti C, and Vogelstein B. Cancer etiology. Variation in cancer risk among tissues
can be explained by the number of stem cell divisions. Science. 2015;347(6217):78-81.
10. Connor AA, Denroche RE, Jang GH, Timms L, Kalimuthu SN, Selander I, et al.
Association of Distinct Mutational Signatures With Correlates of Increased Immune
Activity in Pancreatic Ductal Adenocarcinoma. JAMA Oncol. 2017;3(6):774-83.
11. Waddell N, Pajic M, Patch AM, Chang DK, Kassahn KS, Bailey P, et al. Whole genomes
redefine the mutational landscape of pancreatic cancer. Nature. 2015;518(7540):495-501.
12. Collisson EA, Sadanandam A, Olson P, Gibb WJ, Truitt M, Gu S, et al. Subtypes of
pancreatic ductal adenocarcinoma and their differing responses to therapy. Nat Med.
2011;17(4):500-3.
13. Bailey P, Chang DK, Nones K, Johns AL, Patch AM, Gingras MC, et al. Genomic
analyses identify molecular subtypes of pancreatic cancer. Nature. 2016;531(7592):47-
52.
14. Moffitt RA, Marayati R, Flate EL, Volmar KE, Loeza SG, Hoadley KA, et al. Virtual
microdissection identifies distinct tumor- and stroma-specific subtypes of pancreatic
ductal adenocarcinoma. Nat Genet. 2015;47(10):1168-78.
15. Noll EM, Eisen C, Stenzinger A, Espinet E, Muckenhuber A, Klein C, et al. CYP3A5
mediates basal and acquired therapy resistance in different subtypes of pancreatic ductal
adenocarcinoma. Nat Med. 2016;22(3):278-87.
37
16. Farrell JJ, Elsaleh H, Garcia M, Lai R, Ammar A, Regine WF, et al. Human equilibrative
nucleoside transporter 1 levels predict response to gemcitabine in patients with pancreatic
cancer. Gastroenterology. 2009;136(1):187-95.
17. Jones S, Zhang X, Parsons DW, Lin JC, Leary RJ, Angenendt P, et al. Core signaling
pathways in human pancreatic cancers revealed by global genomic analyses. Science.
2008;321(5897):1801-6.
18. Hanahan D, and Weinberg RA. Hallmarks of cancer: the next generation. Cell.
2011;144(5):646-74.
19. Biankin AV, Waddell N, Kassahn KS, Gingras MC, Muthuswamy LB, Johns AL, et al.
Pancreatic cancer genomes reveal aberrations in axon guidance pathway genes. Nature.
2012;491(7424):399-405.
20. Kaukonen R, Mai A, Georgiadou M, Saari M, De Franceschi N, Betz T, et al. Normal
stroma suppresses cancer cell proliferation via mechanosensitive regulation of JMJD1a-
mediated transcription. Nat Commun. 2016;7:12237.
21. Rath N, and Olson MF. Regulation of pancreatic cancer aggressiveness by stromal
stiffening. Nat Med. 2016;22(5):462-3.
22. Ansari D, Carvajo M, Bauden M, and Andersson R. Pancreatic cancer stroma:
controversies and current insights. Scand J Gastroenterol. 2017;52(6-7):641-6.
23. Neesse A, Algul H, Tuveson DA, and Gress TM. Stromal biology and therapy in
pancreatic cancer: a changing paradigm. Gut. 2015;64(9):1476-84.
24. Lunardi S, Jamieson NB, Lim SY, Griffiths KL, Carvalho-Gaspar M, Al-Assar O, et al.
IP-10/CXCL10 induction in human pancreatic cancer stroma influences lymphocytes
recruitment and correlates with poor survival. Oncotarget. 2014;5(22):11064-80.
25. Ene-Obong A, Clear AJ, Watt J, Wang J, Fatah R, Riches JC, et al. Activated pancreatic
stellate cells sequester CD8+ T cells to reduce their infiltration of the juxtatumoral
compartment of pancreatic ductal adenocarcinoma. Gastroenterology. 2013;145(5):1121-
32.
26. Sherman MH, Yu RT, Engle DD, Ding N, Atkins AR, Tiriac H, et al. Vitamin D
receptor-mediated stromal reprogramming suppresses pancreatitis and enhances
pancreatic cancer therapy. Cell. 2014;159(1):80-93.
27. Watt J, and Kocher HM. The desmoplastic stroma of pancreatic cancer is a barrier to
immune cell infiltration. Oncoimmunology. 2013;2(12):e26788.
28. Poschke I, Faryna M, Bergmann F, Flossdorf M, Lauenstein C, Hermes J, et al.
Identification of a tumor-reactive T-cell repertoire in the immune infiltrate of patients
with resectable pancreatic ductal adenocarcinoma. Oncoimmunology.
2016;5(12):e1240859.
29. Balachandran VP, Luksza M, Zhao JN, Makarov V, Moral JA, Remark R, et al.
Identification of unique neoantigen qualities in long-term survivors of pancreatic cancer.
Nature. 2017;551(7681):512-6.
30. Vonderheide RH, and Bayne LJ. Inflammatory networks and immune surveillance of
pancreatic carcinoma. Curr Opin Immunol. 2013;25(2):200-5.
31. Evans RA, Diamond MS, Rech AJ, Chao T, Richardson MW, Lin JH, et al. Lack of
immunoediting in murine pancreatic cancer reversed with neoantigen. JCI Insight.
2016;1(14).
32. Hwang CI, Boj SF, Clevers H, and Tuveson DA. Preclinical models of pancreatic ductal
adenocarcinoma. J Pathol. 2016;238(2):197-204.
38
33. Sato T, Vries RG, Snippert HJ, van de Wetering M, Barker N, Stange DE, et al. Single
Lgr5 stem cells build crypt-villus structures in vitro without a mesenchymal niche.
Nature. 2009;459(7244):262-5.
34. Boj SF, Hwang CI, Baker LA, Chio, II, Engle DD, Corbo V, et al. Organoid models of
human and mouse ductal pancreatic cancer. Cell. 2015;160(1-2):324-38.
35. Tiriac H, Belleau P, Engle DD, Plenker D, Deschenes A, Somerville TDD, et al.
Organoid Profiling Identifies Common Responders to Chemotherapy in Pancreatic
Cancer. Cancer Discov. 2018;8(9):1112-29.
36. Guo X, Hollander L, MacPherson D, Wang L, Velazquez H, Chang J, et al. Inhibition of
renalase expression and signaling has antitumor activity in pancreatic cancer. Sci Rep.
2016;6:22996.
37. Patra KC, Bardeesy N, and Mizukami Y. Diversity of Precursor Lesions For Pancreatic
Cancer: The Genetics and Biology of Intraductal Papillary Mucinous Neoplasm. Clin
Transl Gastroenterol. 2017;8(4):e86.
VII. FIGURES
Figure 1. Genetic evolution of pancreatic cancer. Pancreatic cancer may arise from either the
development and progression of intraductal papillary mucinous neoplasm (top) or pancreatic
intraepithelial neoplasm (bottom) as a result of sequential accumulation of characteristic driver
mutations. This illustration was adapted from REF 37, with permission.
39
Figure 2. Creation of immunogenic murine organoid models of PDAC using KP-NINJA
mouse model. (A) Schematic representation of major steps involved in the isolation of murine
pancreatic organoids. (B) Genetic features of KP-NINJA mouse model for Cre-recombinase
inducible mutation of KRAS and deletion of P53 (top), and multilayered control of inducible
expression of GFP-tagged T cell neoantigens by Cre recombinase, rtTA-doxycycline and FLPo-
40
tamoxifen (bottom). GFP, green fluorescent protein; rtTA, reverse tetracycline-controlled
transactivator; FLPo, codon-improved flippase recombinase.
Figure 3. In vitro transformed murine pancreatic organoids recapitulates features of early
PDAC in mouse. (A) Neoplastic transformation of murine pancreatic organoids by lentivirus
encoding rtTA-Cre-iRFP670. Fluorescence imaging confirms RFP labeling of transformed cells
in organoids. (B) Flow cytometry analysis confirms RFP expression in transformed organoids,
which were subsequently sorted for the brightest 10% of cells expressing RFP. Leaky expression
41
of GFP-tagged neoantigens is not observed in these organoids. Untransformed organoids were
used as a negative control. (C) Subcutaneous injection of transformed versus untransformed
organoids in opposite flanks of mouse results in tumor formation only with transformed
organoids. (D) H&E of tumor derived from subcutaneous injection transformed organoids in
mouse. (E) H&E of primary pancreatic tumor from KP-C mouse model. (F) H&E of tumor
derived from subcutaneous injection of PDAC cell lines generated from KP-C mouse model.
42
Figure 4. Modeling PDAC progression by serial in vivo transfer of transformed murine
pancreatic organoids. (A) Experimental design for serial in vivo transfer of organoids. (B)
H&E of organoid-derived tumors after successive rounds of in vivo transfer shows progressively
more advanced tumors. (C) Immunohistochemical analysis of organoid-derived tumors after one
round of in vivo transfer. (D) Fluorescence imaging of organoids reveals more uniform RFP
labeling of organoids after one round of in vivo transfer versus organoids before in vivo transfer,
indicating in vivo selection of transformed neoplastic organoids
43
Figure 5. High level of neoantigen expression in murine PDAC organoids results in
rejection of tumor growth in mouse. (A) Experimental design for expression of GFP-tagged
neoantigens in PDAC organoids by adenovirus encoding FLPo. (B) Flow cytometry analysis
confirms GFP expression in organoids treated with adenovirus encoding FLPo, which were
subsequently sorted for the brightest 10% of GFP-positive cells. Organoids treated with
adenovirus encoding Cre was used as a negative control. (C) Subcutaneous injection of
neoantigen-positive versus neoantigen-negative transformed organoids. First cohort of mice
received neoantigen-negative organoids (N=5). Second cohort received neoantigen-positive
organoids (N=6). Third cohort received neoantigen-positive organoids plus retroorbital injections
of luciferase-positive P14 CD8+ T cells 24 hours prior to organoid injections (N=3). In vivo
imaging after 24 hours of organoid injections reveals accumulation of luciferase-positive T cells
at the site of organoid injections in the third cohort. (D) Mice were monitored for growth of
tumors for up to 30 days. Tumor growth was observed in all of the mice in the first cohort. There
was no tumor growth in any mouse in the second or third cohort that received neonantigen-
positive organoids.
44
Figure 6. Low level of neoantigen expression in murine PDAC organoids permits tumor
growth with increased immune infiltration. (A) A murine PDAC organoid line that expresses
GFP-tagged T cell neoantigens at a low level was generated by dilution of neoantigen-positive
organoids with neoantigen-negative organoids. Flow cytometry analysis confirms GFP
expression in only 10% of the total population. (B) Subcutaneous injection of organoids
generated from (A) resulted in growth of tumors in mouse. H&E of tumors derived from these
organoids shows increased immune infiltration compared to tumors derived from neoantigen-
negative organoids.
45
Figure 7. Development of co-culture model system for murine PDAC organoids and T cells.
(A) P14 CD8+ T cells were sorted from splenocytes of P14 mouse following in vitro expansion
and pre-activation with IL-2 and GP33 peptide, respectively. Blue calcein dye was used to label
46
prepared T cells and green calcein dye was used to label both T cells and PDAC organoids. T
cells were introduced to the liquid medium of wells containing either neoantigen-positive or
neoantigen-negative PDAC organoids which were maintained in Matrigel plugs in 24-well plate
format. Fluorescence imaging of co-cultures at 1 hour demonstrates prominent clustering of T
cells at the boundaries of Matrigel plugs. (B) Fluorescence imaging of co-cultures at 24 hours
reveals evidence of increased T cell infiltration of Matrigel plugs containing neoantigen-positive
organoids. Images were converted to black and white for better visualization of blue dye. IL,
interleukin; GP, glycoprotein.
47
Figure 8. Human PDAC organoids form tumors in mouse that are histologically matched to
patient-derived primary tissues. (A) Schematic overview for the creation of patient-derived
organoids using different types of primary tissues, including EUS-FNB specimens, surgical
resection specimens, and tissues from PDX mouse models that were established by implanting a
piece of surgical resection specimen in mouse. (B) Patient-derived organoids validated by Sanger
sequencing of organoids for mutations in KRAS and P53, in vivo transfer of organoids for tumor
48
formation in mouse, and IHC analysis of organoid-derived tumors. (C) IHC analyses of tumors
generated from different types of primary tissues are shown. Row X shows a tumor generated
from organoids derived from FNB. Row Y shows a tumor directly taken from a PDX mouse
model. Row Y’ shows a tumor generated from organoids derived from the tumor shown in row
Y. Tumors shown in rows Y and Y’ were ultimately derived from the same patient and are
histologically matched. (D) H&E of three additional tumors derived from FNB specimens are
shown. Bx120817 (left) was derived from the primary tumor of a patient who had borderline
resectable disease. Bx111417 (middle) was derived from the primary tumor of a patient who had
metastatic disease. Bx102417 (right) was derived from a metastatic lesion in the liver. (E) Sanger
sequencing of patient-derived organoids reveals classic G12V mutation in KRAS. PDX060917
(left) was derived from tissues from a PDX mouse model and Bx011218A (right) was derived
from an EUS-FNB specimen. EUS-FNB, endoscopic ultrasound-guided fine-needle biopsy;
PDX, patient-derived xenograft; IHC, immunohistochemistry.
VIII. TABLES
Mutated
gene
Frequency
(%)
Effect of
mutation
Cellular process or
pathway affected
Biological significance of
mutation
KRAS 95 Gain of
function
RAS–ERK
pathway
Ligand-independent cell
proliferation and survival;
immunosuppression; metabolic
alterations
CDKN2A 90 Loss of
function
G1/S transition G1/S checkpoint failure
TP53 80-85 Gain of
function
DNA damage
response
G1/S checkpoint failure; G2/M
checkpoint failure; apoptosis
resistance
SMAD4 55 Loss of
function
TGFβ pathway Failure of celluar homeostasis;
loss of TGFβ- and TP53-mediated
gene expression
TGFBR1 ≤10 Loss of
function
TGFβ pathway Failure of celluar homeostasis;
loss of TGFβ- and TP53-mediated
gene expression
49
TGFBR2 ≤10 Loss of
function
TGFβ pathway Failure of celluar homeostasis;
loss of TGFβ- and TP53-mediated
gene expression
ARID1A ≤10 Loss of
function
Epigenomic
reprogramming -
SWI/SNF
Loss of regulatory function in
modulating nucleosomal DNA-
histone interactions
ARID1B ≤10 Loss of
function
Epigenomic
reprogramming -
SWI/SNF
Loss of regulatory function in
modulating nucleosomal DNA-
histone interactions
ARID2 ≤10 Loss of
function
Epigenomic
reprogramming -
SWI/SNF
Loss of regulatory function in
modulating nucleosomal DNA-
histone interactions
KMT2C ≤10 Loss of
function
Epigenomic
reprogramming -
KMT2
Decreased methylation of H3K4
KMT2D ≤10 Loss of
function
Epigenomic
reprogramming -
KMT2
Decreased methylation of H3K4
KMT2A ≤10 Loss of
function
Epigenomic
reprogramming -
KMT2
Decreased methylation of H3K4
SF3B1 ≤10 Altered
function
RNA splicing Loss of polycomb repressive
complex-mediated transcriptional
regulation of HOX genes;
abnormal splicing of pre-mRNA
PCDH15 ≤10 Loss of
function
Homophilic cell
adhesion
Disruption of cadherin-mediated
calcium-dependent cell adhesion
BRAF ≤5 Gain of
function
RAS–ERK
pathway
Ligand-independent cell
proliferation and survival;
immunosuppression; metabolic
alterations
APC2 ≤5 Loss of
function
G1/S transition G1/S checkpoint failure
CHD1 ≤5 Loss of
function
G1/S transition G1/S checkpoint failure
FBXW7 ≤5 Loss of
function
G1/S transition G1/S checkpoint failure
ATM ≤5 Loss of
function
DNA damage
response
G1/S checkpoint failure; G2/M
checkpoint failure; apoptosis
resistance
ACVR1B ≤5 Loss of
function
TGFβ pathway Failure of celluar homeostasis;
loss of TGFβ- and TP53-mediated
gene expression
SMAD3 ≤5 Loss of
function
TGFβ pathway Failure of celluar homeostasis;
loss of TGFβ- and TP53-mediated
gene expression
50
PBRM1 ≤5 Loss of
function
Epigenomic
reprogramming -
SWI/SNF
Loss of regulatory function in
modulating nucleosomal DNA-
histone interactions
SMARCA2 ≤5 Loss of
function
Epigenomic
reprogramming -
SWI/SNF
Loss of regulatory function in
modulating nucleosomal DNA-
histone interactions
SMARCA4 ≤5 Loss of
function
Epigenomic
reprogramming -
SWI/SNF
Loss of regulatory function in
modulating nucleosomal DNA-
histone interactions
MKK4 ≤5 Loss of
function
Cellular stress
response
Failure of JNK signaling;
disruption of TLR signaling
ROBO1 ≤5 Loss of
function
Axon guidance Abnormal migration of cells
ROBO2 ≤5 Loss of
function
Axon guidance Abnormal migration of cells
SLIT ≤5 Loss of
function
Axon guidance Abnormal migration of cells
Table 1. Mutational landscape of pancreatic cancer. Commonly mutated genes in PDAC are
organized by frequency of mutation in PDAC, effect of mutation on gene function, celluar
process or signaling pathway affected, and biological significance of mutation. This table is a
summary of data described in greater detail in REF 6.

More Related Content

What's hot

JLS-064-077-MASTANEH-ABNORMAL-PATIENTS(1)
JLS-064-077-MASTANEH-ABNORMAL-PATIENTS(1)JLS-064-077-MASTANEH-ABNORMAL-PATIENTS(1)
JLS-064-077-MASTANEH-ABNORMAL-PATIENTS(1)mastaneh zohri
 
The application of extracorporeal photochemotherapy to head and neck squamous...
The application of extracorporeal photochemotherapy to head and neck squamous...The application of extracorporeal photochemotherapy to head and neck squamous...
The application of extracorporeal photochemotherapy to head and neck squamous...TÀI LIỆU NGÀNH MAY
 
Genes and Tissue Culture Assignment Presentation (Group 3)
Genes and Tissue Culture Assignment Presentation (Group 3)Genes and Tissue Culture Assignment Presentation (Group 3)
Genes and Tissue Culture Assignment Presentation (Group 3)Lim Ke Wen
 
Kim Solez TEP meets Human Cell Atlas a glimpse into future of pathology winte...
Kim Solez TEP meets Human Cell Atlas a glimpse into future of pathology winte...Kim Solez TEP meets Human Cell Atlas a glimpse into future of pathology winte...
Kim Solez TEP meets Human Cell Atlas a glimpse into future of pathology winte...Kim Solez ,
 
SCT60103 Group 4 Assignment
SCT60103 Group 4 AssignmentSCT60103 Group 4 Assignment
SCT60103 Group 4 AssignmentSheryn Yeo
 
Is Cancer a Genetic Disease? | The Cancer Genome Atlas Project Results
Is Cancer a Genetic Disease? | The Cancer Genome Atlas Project ResultsIs Cancer a Genetic Disease? | The Cancer Genome Atlas Project Results
Is Cancer a Genetic Disease? | The Cancer Genome Atlas Project ResultsMarkSloan21
 
Mike (Gang) CV-updated
Mike (Gang) CV-updatedMike (Gang) CV-updated
Mike (Gang) CV-updatedGang Zhang
 
Animal Research Models:Potential
Animal Research Models:PotentialAnimal Research Models:Potential
Animal Research Models:PotentialAsra Nasir Khan
 
Informatics and data analytics to support for exposome-based discovery
Informatics and data analytics to support for exposome-based discoveryInformatics and data analytics to support for exposome-based discovery
Informatics and data analytics to support for exposome-based discoveryChirag Patel
 
PTG-E Congress 2008 Microbe Hunters GI Lecture
PTG-E Congress 2008 Microbe Hunters GI LecturePTG-E Congress 2008 Microbe Hunters GI Lecture
PTG-E Congress 2008 Microbe Hunters GI Lecturemarlicz
 
Intro to Biomedical Informatics 701
Intro to Biomedical Informatics 701 Intro to Biomedical Informatics 701
Intro to Biomedical Informatics 701 Chirag Patel
 
Big data and the exposome, Oregon State 040616
Big data and the exposome, Oregon State 040616Big data and the exposome, Oregon State 040616
Big data and the exposome, Oregon State 040616Chirag Patel
 
Research associate resume
Research associate resumeResearch associate resume
Research associate resumeShravida Shetty
 
GROUP 7- KNOCK IN MOUSE MODEL USED FOR DISEASE MODELLING
GROUP 7- KNOCK IN MOUSE MODEL USED FOR DISEASE MODELLINGGROUP 7- KNOCK IN MOUSE MODEL USED FOR DISEASE MODELLING
GROUP 7- KNOCK IN MOUSE MODEL USED FOR DISEASE MODELLINGVinitha Govindan Rajan
 
Improvisation of Conventional Techniques: The Future of Oncology Research
Improvisation of Conventional Techniques: The Future of Oncology ResearchImprovisation of Conventional Techniques: The Future of Oncology Research
Improvisation of Conventional Techniques: The Future of Oncology Researchasclepiuspdfs
 
Ethics And The Human Genome Project
Ethics And The Human Genome ProjectEthics And The Human Genome Project
Ethics And The Human Genome ProjectNadia Blanchard
 

What's hot (20)

JLS-064-077-MASTANEH-ABNORMAL-PATIENTS(1)
JLS-064-077-MASTANEH-ABNORMAL-PATIENTS(1)JLS-064-077-MASTANEH-ABNORMAL-PATIENTS(1)
JLS-064-077-MASTANEH-ABNORMAL-PATIENTS(1)
 
The application of extracorporeal photochemotherapy to head and neck squamous...
The application of extracorporeal photochemotherapy to head and neck squamous...The application of extracorporeal photochemotherapy to head and neck squamous...
The application of extracorporeal photochemotherapy to head and neck squamous...
 
Genes and Tissue Culture Assignment Presentation (Group 3)
Genes and Tissue Culture Assignment Presentation (Group 3)Genes and Tissue Culture Assignment Presentation (Group 3)
Genes and Tissue Culture Assignment Presentation (Group 3)
 
Kim Solez TEP meets Human Cell Atlas a glimpse into future of pathology winte...
Kim Solez TEP meets Human Cell Atlas a glimpse into future of pathology winte...Kim Solez TEP meets Human Cell Atlas a glimpse into future of pathology winte...
Kim Solez TEP meets Human Cell Atlas a glimpse into future of pathology winte...
 
SCT60103 Group 4 Assignment
SCT60103 Group 4 AssignmentSCT60103 Group 4 Assignment
SCT60103 Group 4 Assignment
 
Is Cancer a Genetic Disease? | The Cancer Genome Atlas Project Results
Is Cancer a Genetic Disease? | The Cancer Genome Atlas Project ResultsIs Cancer a Genetic Disease? | The Cancer Genome Atlas Project Results
Is Cancer a Genetic Disease? | The Cancer Genome Atlas Project Results
 
Micro
MicroMicro
Micro
 
Biomarker-Vol-8
Biomarker-Vol-8Biomarker-Vol-8
Biomarker-Vol-8
 
Mike (Gang) CV-updated
Mike (Gang) CV-updatedMike (Gang) CV-updated
Mike (Gang) CV-updated
 
Animal Research Models:Potential
Animal Research Models:PotentialAnimal Research Models:Potential
Animal Research Models:Potential
 
Informatics and data analytics to support for exposome-based discovery
Informatics and data analytics to support for exposome-based discoveryInformatics and data analytics to support for exposome-based discovery
Informatics and data analytics to support for exposome-based discovery
 
PTG-E Congress 2008 Microbe Hunters GI Lecture
PTG-E Congress 2008 Microbe Hunters GI LecturePTG-E Congress 2008 Microbe Hunters GI Lecture
PTG-E Congress 2008 Microbe Hunters GI Lecture
 
Us Patent US9415087
Us Patent US9415087  Us Patent US9415087
Us Patent US9415087
 
Intro to Biomedical Informatics 701
Intro to Biomedical Informatics 701 Intro to Biomedical Informatics 701
Intro to Biomedical Informatics 701
 
Big data and the exposome, Oregon State 040616
Big data and the exposome, Oregon State 040616Big data and the exposome, Oregon State 040616
Big data and the exposome, Oregon State 040616
 
Research associate resume
Research associate resumeResearch associate resume
Research associate resume
 
GROUP 7- KNOCK IN MOUSE MODEL USED FOR DISEASE MODELLING
GROUP 7- KNOCK IN MOUSE MODEL USED FOR DISEASE MODELLINGGROUP 7- KNOCK IN MOUSE MODEL USED FOR DISEASE MODELLING
GROUP 7- KNOCK IN MOUSE MODEL USED FOR DISEASE MODELLING
 
Improvisation of Conventional Techniques: The Future of Oncology Research
Improvisation of Conventional Techniques: The Future of Oncology ResearchImprovisation of Conventional Techniques: The Future of Oncology Research
Improvisation of Conventional Techniques: The Future of Oncology Research
 
s12935-014-0115-7
s12935-014-0115-7s12935-014-0115-7
s12935-014-0115-7
 
Ethics And The Human Genome Project
Ethics And The Human Genome ProjectEthics And The Human Genome Project
Ethics And The Human Genome Project
 

Similar to Development of Pancreatic Cancer Organoid Models to Study Immune Response

Organoid culture in cancer
Organoid culture in cancerOrganoid culture in cancer
Organoid culture in cancerSamieh Asadian
 
Kholood Ahmad Poster FINAL 4 17 2015
Kholood Ahmad Poster FINAL 4 17 2015Kholood Ahmad Poster FINAL 4 17 2015
Kholood Ahmad Poster FINAL 4 17 2015Kholood Ahmad
 
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...semualkaira
 
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...semualkaira
 
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...semualkaira
 
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...semualkaira
 
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...semualkaira
 
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...semualkaira
 
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...semualkaira
 
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...semualkaira
 
The role of traf3 and cyld mutationin the etiology of human papillomavirus dr...
The role of traf3 and cyld mutationin the etiology of human papillomavirus dr...The role of traf3 and cyld mutationin the etiology of human papillomavirus dr...
The role of traf3 and cyld mutationin the etiology of human papillomavirus dr...TÀI LIỆU NGÀNH MAY
 
Collision Tumour-IRB -Final.pptx
Collision Tumour-IRB -Final.pptxCollision Tumour-IRB -Final.pptx
Collision Tumour-IRB -Final.pptxsherin110346
 
Deriving Mesenchymal Stem Cells from Human Amniotic Fluid – Potential for an ...
Deriving Mesenchymal Stem Cells from Human Amniotic Fluid – Potential for an ...Deriving Mesenchymal Stem Cells from Human Amniotic Fluid – Potential for an ...
Deriving Mesenchymal Stem Cells from Human Amniotic Fluid – Potential for an ...cordbloodsymposium
 
Advances in experimental medicine and biology hussain book
Advances in experimental medicine and biology hussain bookAdvances in experimental medicine and biology hussain book
Advances in experimental medicine and biology hussain bookmantu verma
 
Ultrasound technique offers new approach to fighting cancer
Ultrasound technique offers new approach to fighting cancerUltrasound technique offers new approach to fighting cancer
Ultrasound technique offers new approach to fighting cancerAtlantis Worldwide LLC
 
Pharmacological screening by harikesh maurya
Pharmacological screening by harikesh mauryaPharmacological screening by harikesh maurya
Pharmacological screening by harikesh mauryaHarikesh Maurya
 
Uncovering intratumoral and intertumoral heterogeneity among single cell canc...
Uncovering intratumoral and intertumoral heterogeneity among single cell canc...Uncovering intratumoral and intertumoral heterogeneity among single cell canc...
Uncovering intratumoral and intertumoral heterogeneity among single cell canc...https://www.facebook.com/garmentspace
 
1 s2.0-s0015028214018603-main
1 s2.0-s0015028214018603-main1 s2.0-s0015028214018603-main
1 s2.0-s0015028214018603-main鋒博 蔡
 

Similar to Development of Pancreatic Cancer Organoid Models to Study Immune Response (20)

Organoid culture in cancer
Organoid culture in cancerOrganoid culture in cancer
Organoid culture in cancer
 
1.g.2014-patho~ (1.introduction-wyl)
  1.g.2014-patho~ (1.introduction-wyl)  1.g.2014-patho~ (1.introduction-wyl)
1.g.2014-patho~ (1.introduction-wyl)
 
Kholood Ahmad Poster FINAL 4 17 2015
Kholood Ahmad Poster FINAL 4 17 2015Kholood Ahmad Poster FINAL 4 17 2015
Kholood Ahmad Poster FINAL 4 17 2015
 
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
 
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
 
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
 
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
 
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
 
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
 
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
 
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
Preliminary Study on Monitoring Drug Resistance of Colon Cancer with Intravox...
 
The role of traf3 and cyld mutationin the etiology of human papillomavirus dr...
The role of traf3 and cyld mutationin the etiology of human papillomavirus dr...The role of traf3 and cyld mutationin the etiology of human papillomavirus dr...
The role of traf3 and cyld mutationin the etiology of human papillomavirus dr...
 
Collision Tumour-IRB -Final.pptx
Collision Tumour-IRB -Final.pptxCollision Tumour-IRB -Final.pptx
Collision Tumour-IRB -Final.pptx
 
Deriving Mesenchymal Stem Cells from Human Amniotic Fluid – Potential for an ...
Deriving Mesenchymal Stem Cells from Human Amniotic Fluid – Potential for an ...Deriving Mesenchymal Stem Cells from Human Amniotic Fluid – Potential for an ...
Deriving Mesenchymal Stem Cells from Human Amniotic Fluid – Potential for an ...
 
Advances in experimental medicine and biology hussain book
Advances in experimental medicine and biology hussain bookAdvances in experimental medicine and biology hussain book
Advances in experimental medicine and biology hussain book
 
uts-vol1
uts-vol1uts-vol1
uts-vol1
 
Ultrasound technique offers new approach to fighting cancer
Ultrasound technique offers new approach to fighting cancerUltrasound technique offers new approach to fighting cancer
Ultrasound technique offers new approach to fighting cancer
 
Pharmacological screening by harikesh maurya
Pharmacological screening by harikesh mauryaPharmacological screening by harikesh maurya
Pharmacological screening by harikesh maurya
 
Uncovering intratumoral and intertumoral heterogeneity among single cell canc...
Uncovering intratumoral and intertumoral heterogeneity among single cell canc...Uncovering intratumoral and intertumoral heterogeneity among single cell canc...
Uncovering intratumoral and intertumoral heterogeneity among single cell canc...
 
1 s2.0-s0015028214018603-main
1 s2.0-s0015028214018603-main1 s2.0-s0015028214018603-main
1 s2.0-s0015028214018603-main
 

Recently uploaded

Solving Puzzles Benefits Everyone (English).pptx
Solving Puzzles Benefits Everyone (English).pptxSolving Puzzles Benefits Everyone (English).pptx
Solving Puzzles Benefits Everyone (English).pptxOH TEIK BIN
 
Sanyam Choudhary Chemistry practical.pdf
Sanyam Choudhary Chemistry practical.pdfSanyam Choudhary Chemistry practical.pdf
Sanyam Choudhary Chemistry practical.pdfsanyamsingh5019
 
Interactive Powerpoint_How to Master effective communication
Interactive Powerpoint_How to Master effective communicationInteractive Powerpoint_How to Master effective communication
Interactive Powerpoint_How to Master effective communicationnomboosow
 
_Math 4-Q4 Week 5.pptx Steps in Collecting Data
_Math 4-Q4 Week 5.pptx Steps in Collecting Data_Math 4-Q4 Week 5.pptx Steps in Collecting Data
_Math 4-Q4 Week 5.pptx Steps in Collecting DataJhengPantaleon
 
“Oh GOSH! Reflecting on Hackteria's Collaborative Practices in a Global Do-It...
“Oh GOSH! Reflecting on Hackteria's Collaborative Practices in a Global Do-It...“Oh GOSH! Reflecting on Hackteria's Collaborative Practices in a Global Do-It...
“Oh GOSH! Reflecting on Hackteria's Collaborative Practices in a Global Do-It...Marc Dusseiller Dusjagr
 
Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...
Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...
Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...EduSkills OECD
 
CARE OF CHILD IN INCUBATOR..........pptx
CARE OF CHILD IN INCUBATOR..........pptxCARE OF CHILD IN INCUBATOR..........pptx
CARE OF CHILD IN INCUBATOR..........pptxGaneshChakor2
 
How to Make a Pirate ship Primary Education.pptx
How to Make a Pirate ship Primary Education.pptxHow to Make a Pirate ship Primary Education.pptx
How to Make a Pirate ship Primary Education.pptxmanuelaromero2013
 
Accessible design: Minimum effort, maximum impact
Accessible design: Minimum effort, maximum impactAccessible design: Minimum effort, maximum impact
Accessible design: Minimum effort, maximum impactdawncurless
 
Kisan Call Centre - To harness potential of ICT in Agriculture by answer farm...
Kisan Call Centre - To harness potential of ICT in Agriculture by answer farm...Kisan Call Centre - To harness potential of ICT in Agriculture by answer farm...
Kisan Call Centre - To harness potential of ICT in Agriculture by answer farm...Krashi Coaching
 
The basics of sentences session 2pptx copy.pptx
The basics of sentences session 2pptx copy.pptxThe basics of sentences session 2pptx copy.pptx
The basics of sentences session 2pptx copy.pptxheathfieldcps1
 
URLs and Routing in the Odoo 17 Website App
URLs and Routing in the Odoo 17 Website AppURLs and Routing in the Odoo 17 Website App
URLs and Routing in the Odoo 17 Website AppCeline George
 
Alper Gobel In Media Res Media Component
Alper Gobel In Media Res Media ComponentAlper Gobel In Media Res Media Component
Alper Gobel In Media Res Media ComponentInMediaRes1
 
Mastering the Unannounced Regulatory Inspection
Mastering the Unannounced Regulatory InspectionMastering the Unannounced Regulatory Inspection
Mastering the Unannounced Regulatory InspectionSafetyChain Software
 
Science 7 - LAND and SEA BREEZE and its Characteristics
Science 7 - LAND and SEA BREEZE and its CharacteristicsScience 7 - LAND and SEA BREEZE and its Characteristics
Science 7 - LAND and SEA BREEZE and its CharacteristicsKarinaGenton
 
Contemporary philippine arts from the regions_PPT_Module_12 [Autosaved] (1).pptx
Contemporary philippine arts from the regions_PPT_Module_12 [Autosaved] (1).pptxContemporary philippine arts from the regions_PPT_Module_12 [Autosaved] (1).pptx
Contemporary philippine arts from the regions_PPT_Module_12 [Autosaved] (1).pptxRoyAbrique
 
Concept of Vouching. B.Com(Hons) /B.Compdf
Concept of Vouching. B.Com(Hons) /B.CompdfConcept of Vouching. B.Com(Hons) /B.Compdf
Concept of Vouching. B.Com(Hons) /B.CompdfUmakantAnnand
 

Recently uploaded (20)

Solving Puzzles Benefits Everyone (English).pptx
Solving Puzzles Benefits Everyone (English).pptxSolving Puzzles Benefits Everyone (English).pptx
Solving Puzzles Benefits Everyone (English).pptx
 
Model Call Girl in Tilak Nagar Delhi reach out to us at 🔝9953056974🔝
Model Call Girl in Tilak Nagar Delhi reach out to us at 🔝9953056974🔝Model Call Girl in Tilak Nagar Delhi reach out to us at 🔝9953056974🔝
Model Call Girl in Tilak Nagar Delhi reach out to us at 🔝9953056974🔝
 
Sanyam Choudhary Chemistry practical.pdf
Sanyam Choudhary Chemistry practical.pdfSanyam Choudhary Chemistry practical.pdf
Sanyam Choudhary Chemistry practical.pdf
 
Interactive Powerpoint_How to Master effective communication
Interactive Powerpoint_How to Master effective communicationInteractive Powerpoint_How to Master effective communication
Interactive Powerpoint_How to Master effective communication
 
_Math 4-Q4 Week 5.pptx Steps in Collecting Data
_Math 4-Q4 Week 5.pptx Steps in Collecting Data_Math 4-Q4 Week 5.pptx Steps in Collecting Data
_Math 4-Q4 Week 5.pptx Steps in Collecting Data
 
“Oh GOSH! Reflecting on Hackteria's Collaborative Practices in a Global Do-It...
“Oh GOSH! Reflecting on Hackteria's Collaborative Practices in a Global Do-It...“Oh GOSH! Reflecting on Hackteria's Collaborative Practices in a Global Do-It...
“Oh GOSH! Reflecting on Hackteria's Collaborative Practices in a Global Do-It...
 
Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...
Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...
Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...
 
CARE OF CHILD IN INCUBATOR..........pptx
CARE OF CHILD IN INCUBATOR..........pptxCARE OF CHILD IN INCUBATOR..........pptx
CARE OF CHILD IN INCUBATOR..........pptx
 
9953330565 Low Rate Call Girls In Rohini Delhi NCR
9953330565 Low Rate Call Girls In Rohini  Delhi NCR9953330565 Low Rate Call Girls In Rohini  Delhi NCR
9953330565 Low Rate Call Girls In Rohini Delhi NCR
 
How to Make a Pirate ship Primary Education.pptx
How to Make a Pirate ship Primary Education.pptxHow to Make a Pirate ship Primary Education.pptx
How to Make a Pirate ship Primary Education.pptx
 
TataKelola dan KamSiber Kecerdasan Buatan v022.pdf
TataKelola dan KamSiber Kecerdasan Buatan v022.pdfTataKelola dan KamSiber Kecerdasan Buatan v022.pdf
TataKelola dan KamSiber Kecerdasan Buatan v022.pdf
 
Accessible design: Minimum effort, maximum impact
Accessible design: Minimum effort, maximum impactAccessible design: Minimum effort, maximum impact
Accessible design: Minimum effort, maximum impact
 
Kisan Call Centre - To harness potential of ICT in Agriculture by answer farm...
Kisan Call Centre - To harness potential of ICT in Agriculture by answer farm...Kisan Call Centre - To harness potential of ICT in Agriculture by answer farm...
Kisan Call Centre - To harness potential of ICT in Agriculture by answer farm...
 
The basics of sentences session 2pptx copy.pptx
The basics of sentences session 2pptx copy.pptxThe basics of sentences session 2pptx copy.pptx
The basics of sentences session 2pptx copy.pptx
 
URLs and Routing in the Odoo 17 Website App
URLs and Routing in the Odoo 17 Website AppURLs and Routing in the Odoo 17 Website App
URLs and Routing in the Odoo 17 Website App
 
Alper Gobel In Media Res Media Component
Alper Gobel In Media Res Media ComponentAlper Gobel In Media Res Media Component
Alper Gobel In Media Res Media Component
 
Mastering the Unannounced Regulatory Inspection
Mastering the Unannounced Regulatory InspectionMastering the Unannounced Regulatory Inspection
Mastering the Unannounced Regulatory Inspection
 
Science 7 - LAND and SEA BREEZE and its Characteristics
Science 7 - LAND and SEA BREEZE and its CharacteristicsScience 7 - LAND and SEA BREEZE and its Characteristics
Science 7 - LAND and SEA BREEZE and its Characteristics
 
Contemporary philippine arts from the regions_PPT_Module_12 [Autosaved] (1).pptx
Contemporary philippine arts from the regions_PPT_Module_12 [Autosaved] (1).pptxContemporary philippine arts from the regions_PPT_Module_12 [Autosaved] (1).pptx
Contemporary philippine arts from the regions_PPT_Module_12 [Autosaved] (1).pptx
 
Concept of Vouching. B.Com(Hons) /B.Compdf
Concept of Vouching. B.Com(Hons) /B.CompdfConcept of Vouching. B.Com(Hons) /B.Compdf
Concept of Vouching. B.Com(Hons) /B.Compdf
 

Development of Pancreatic Cancer Organoid Models to Study Immune Response

  • 1. Yale University EliScholar – A Digital Platform for Scholarly Publishing at Yale Yale Medicine Thesis Digital Library School of Medicine January 2019 Development Of Pancreatic Cancer Organoid Model For Studying Immune Response In Pancreatic Cancer Jin Woo Yoo Follow this and additional works at: https://elischolar.library.yale.edu/ymtdl This Open Access Thesis is brought to you for free and open access by the School of Medicine at EliScholar – A Digital Platform for Scholarly Publishing at Yale. It has been accepted for inclusion in Yale Medicine Thesis Digital Library by an authorized administrator of EliScholar – A Digital Platform for Scholarly Publishing at Yale. For more information, please contact elischolar@yale.edu. Recommended Citation Yoo, Jin Woo, "Development Of Pancreatic Cancer Organoid Model For Studying Immune Response In Pancreatic Cancer" (2019). Yale Medicine Thesis Digital Library. 3543. https://elischolar.library.yale.edu/ymtdl/3543
  • 2. Development of Pancreatic Cancer Organoid Models for Studying Immune Response in Pancreatic Cancer A Thesis Submitted to the Yale University School of Medicine in Partial Fulfillment of the Requirements for the Degree of Doctor of Medicine by Jin Woo Yoo 2019
  • 3. DEVELOPMENT OF PANCREATIC CANCER ORGANOID MODEL FOR STUDYING IMMUNE RESPONSE IN PANCREATIC CANCER. Jin Woo Yoo, Prashanth R. Gokare, Yevgeniya Foster, Brittany Fitzgerald, Nikhil S. Joshi, James J. Farrell. Section of Gastroenterology, Department of Internal Medicine, Yale University, School of Medicine, New Haven, CT. The importance of immune system in pancreatic ductal adenocarcinoma (PDAC) pathogenesis and therapy remains poorly understood largely due to the lack of effective model systems. Cell lines are not physiologic as they cannot recapitulate the cancer stroma and lose genetic heterogeneity over time. Genetically engineered mouse models of PDAC are more physiologic than cell lines but lack neoantigens needed to mount T cell responses against tumor. Organoid models of PDAC offer unique opportunity to study immune mechanisms in PDAC since organoids can model complex layering of multiple cell types, creating a physiologically relevant system that is highly tractable for genetic manipulation, co-cultures, and high throughput assays. In this study, we sought to establish murine and human organoid models of PDAC to investigate the biology of PDAC immune response, with the specific aims of developing transplantable immunogenic murine PDAC organoid models for the study of antigen- specific anti-tumor T cell responses and assembling a library of experimentally validated, patient-derived PDAC organoid lines for pancreatic cancer precision medicine research. To generate immunogenic murine organoid models of PDAC, pancreatic organoids were isolated from “KP-NINJA” (KrasLox-STOP-Lox-G12D ; P53flox/flox ; inversion induced joined neoantigen) mouse model that has been genetically engineered to express GFP-tagged T cell neoantigens derived from lymphocytic choriomeningitis virus in an inducible fashion. Isolated organoids were transformed in vitro using a lentiviral construct encoding Cre recombinase and
  • 4. RFP reporter for expression of oncogenic KRAS and deletion of P53. A subset of transformed organoids was additionally treated with an adenoviral construct encoding FLPo recombinase to turn on neoantigen expression. Transformed organoids were combined with T cells in both in vivo and in vitro setting to assess for impact on tumor growth. Patient-derived PDAC organoids were generated using endoscopic ultrasound-guided fine needle biopsy (EUS-FNB) specimens, surgical resection specimens, and tissues from patient-derived xenograft mouse models of PDAC. Established human organoid lines were validated by Sanger sequencing, tumor formation in vivo and immunohistochemistry of organoid-derived tumors. Subcutaneous injection of transformed murine PDAC organoids formed tumors in mouse that are histologically similar to early lesions found in human PDAC. Serial in vivo transfer of these organoids by performing sequential rounds of organoid generation from tumors derived from organoids formed progressively more advanced tumors. High level of neoantigen expression in 100% of cells comprising murine PDAC organoids resulted in rejection of tumor growth in mouse, while a low level of neoantigen expression restricted to 10% of cells permitted tumor growth with increased immune infiltration. Expression of neoantigens in T cell-PDAC organoid co-culture model systems promoted T cell infiltration of basement membrane matrix. Additionally, we generated 30+ patient-derived PDAC organoid lines using EUS-FNB and surgical specimens at Yale from 10/2017 to 5/2018. We have successfully established murine and human organoid models of PDAC from various tissues capturing discrete stages of PDAC progression. Our murine organoid models are uniquely equipped to study antigen-specific T cell responses against tumor. Ongoing work includes using CRISPR/Cas9-based lentiviral systems to define genes that impact anti-tumor T cell responses and using patient-derived organoids for precision medicine research.
  • 5. ACKNOWLEDGEMENTS Work for this thesis was completed in the Joshi laboratory under the co-mentorship of James J. Farrell, MD and Nikhil S. Joshi, PhD. Both Dr. Farrell and Dr. Joshi suggested experiments and supervised the work done. Dr. Joshi developed the KP-NINJA mouse model that was fundamental for the creation of immunogenic murine PDAC organoid models. Dr. Farrell performed and provided all the endoscopic ultrasound-guided fine-needle biopsies for the creation of patient-derived PDAC organoid lines. Prashanth Gokare, PhD collaborated with the author on the development of three-dimensional co-culture system for murine pancreatic cancer organoids and T cells and the creation of patient-derived PDAC organoids from surgical resection specimens and their sequencing. Yevgeniya Foster, MD collaborated with the author on creation of immunogenic murine PDAC organoid lines for characterizing immune responses in vivo and immunohistochemical analysis of murine organoid-derived tumors. Brittany Fitzgerald established the primary murine pancreatic cancer cell lines from KP-C mouse and collaborated with the author on in vivo transfer of P14 mouse splenocytes and in vivo imaging for luciferase detection. Marie Robert, MD provided surgical resection specimens for creation of patient- derived pancreatic cancer organoids and interpretation of tumor histology. Ryan Sowell, PhD from Kaech laboratory created the patient-derived xenograft mouse models, some of which were used as source material for the creation of human PDAC organoids. All other experiments were performed independently by the author. Dr. Farrell and Dr. Joshi reviewed and provided comments on the manuscript. National Institute of Health-National Institute of Diabetes and Digestive and Kidney Diseases Medical Student Research Fellowship (T35 grant), Yale University School of Medicine Research Fellowship, and Richard Alan HirshField Memorial Fellowship provided funding to support this work.
  • 6. TABLE OF CONTENTS LIST OF ABBREVIATIONS.......................................................................................................1 INTRODUCTION.........................................................................................................................2 • Background • Cell of Origin • Genetic Landscape of Pancreatic Cancer • Precursor Lesions • Mutational Processes • Tumoral Heterogeneity • Molecular Subtyping of Pancreatic Cancer • Deranged Signaling Pathways / Molecular Aberrations • Tumor Microenvironment • Metabolic Reprogramming • Immune Response in Pancreatic Cancer is Unclear • Pre-clinical Modeling of Pancreatic Cancer STATEMENT OF PURPOSE....................................................................................................18 METHODS...................................................................................................................................19 • Acquisition of human specimens • Isolation and culture of murine pancreatic organoids • Isolation and culture of human PDAC organoids • Isolation of primary murine PDAC cell lines • Genetic manipulation of murine pancreatic organoids • In vivo mouse assays • Immunohistochemical analysis of tumors • Sanger sequencing of organoids • Development of organoid-T cell co-culture model systems RESULTS.....................................................................................................................................26 • KP-NINJA mouse model provides substrate for creation of immunogenic murine organoid models of PDAC • In vitro transformed murine pancreatic organoids form tumors that are histologically similar to early lesions found in human PDAC • Serial in vivo transfer of transformed murine pancreatic organoids results in progressively more advanced tumors • Expression of neoantigens in murine PDAC organoids elicits effective immune response in mouse • Expression of neoantigens in murine PDAC organoids promotes T cell infiltration in T cell-organoid co-culture model • Assembly of human PDAC organoid library
  • 7. 1 DISCUSSION...............................................................................................................................33 REFERENCES.............................................................................................................................36 FIGURES......................................................................................................................................38 TABLES........................................................................................................................................48 LIST OF ABBREVIATIONS CTGF Connective tissue growth factor EGF Epidermal growth factor ER Estrogen receptor ETC Electron transport chain EUS-FNB Endoscopic ultrasound-guided fine needle biopsy FGF Fibroblast growth factor FLP Flippase FRT Flippase recognition target GFP Green fluorescent protein GM-CSF Granulocyte-macrophage colony-stimulating factor GP Glycoprotein hENT1 Human equilibrative nucleoside transporter HGF Hepatocyte growth factor HIF1α Hypoxia-inducible transcription factor 1α HR Homologous recombination IFN-γ Interferon-γ IGF1 Insulin-like growth factor 1 IHC Immunohistochemistry IL-1 Interleukin-1 IPMN Intraductal papillary mucinous neoplasm LCMV Lymphocytic choriomeningitis virus MCN Mucinous cystic neoplasm MDSC Myeloid-derived suppressor cell MMP Matrix metalloproteinase MMR Mismatch repair NF-κB Nuclear factor-κB NSG NOD scid gamma PanIN Pancreatic intraepithelial neoplasm PARP Poly ADP-ribose polymerase PDAC Pancreatic ductal adenocarcinoma PDGF Platelet-derived growth factor PDX Patient-derived xenograft PSC Pancreatic stellate cell
  • 8. 2 RFP Red fluorescent protein rtTA Reverse tetracycline-controlled transactivator STAT3 Signal transducer and activator of transcription 3 TCA Tricyclic acid TCR T cell receptor TGFα Transforming growth factor-α TIMP Tissue inhibitor of metalloproteinases TNFα Tumor necrosis factor-α TRE Tetracycline response element TSLP Thymic stromal lymphopoietin VEGF Vascular endothelial growth factor I. INTRODUCTION Background Pancreatic ductal adenocarcinoma (PDAC; used interchangeably with pancreatic cancer hereafter), the predominant form of pancreatic malignancy, is currently the fourth leading cause of all cancer-related deaths in developed countries and is projected to become second only to lung cancer by year 2024.(1) In 2015 worldwide, 367,000 patients were newly diagnosed with pancreatic cancer, of whom 359,000 patients died due to pancreatic cancer-related causes within the same year.(2) Although surgical resection is currently the only curative treatment for pancreatic cancer, fewer than 20% of patients have resectable disease by the time their diagnosis is made. The overall survival rate at 5 years is less than 7%, with most of the survivors at 5 years belonging to the group of 10-20% of patients who undergo surgical resection of their tumors.(3) Even for those patients undergoing surgery, 80% of them eventually relapse and die from pancreatic cancer. The exceptionally poor prognosis of pancreatic cancer can be attributed to several factors.(2) First is its late diagnosis due to poor early detection, which is delayed by the absence of clear or disease-specific symptoms and the lack of reliable biomarkers for effective screening. Secondly, pancreatic cancer takes an aggressive course, with perineural and vascular invasions
  • 9. 3 and early distant metastases precluding a potentially curative surgical resection. Thirdly, pancreatic cancer displays remarkable resistance to conventional modalities of cancer therapy, including chemotherapy, radiotherapy as well as more recently developed molecularly targeted therapies including immunotherapy. Finally, pancreatic cancer harbors complex tumor biology with both intertumoral and intratumoral genetic heterogeneity, resulting in variable treatment responses from patient to patient thus rendering a generalized approach to therapy difficult. A comprehensive, mechanistic understanding of the pathophysiology underlying pancreatic cancer is fundamental to overcoming these barriers. Cell of Origin The normal pancreas consists of two distinct functional components: endocrine and exocrine. The endocrine component consists of glucagon-producing alpha cells and insulin- producing beta cells that are anatomically organized into islets, and can give rise to a relatively rarer form of pancreatic malignancies termed pancreatic neuroendocrine tumors, which have been found to harbor mutational signatures clearly distinct from those of PDAC. These signatures include inactivation of genes MEN1, ATRX and DAXX, derangements in the mTOR signaling pathway, recurrent YY1 Thr372Arg missense mutations, and biallelic MUTYH inactivating mutations.(4) The exocrine component of the pancreas consists of digestive enzyme-secreting acinar cells and bicarbonate-secreting ductal cells. Historically, ductal cells were thought to be the unique source of PDAC, given their co-expression of epithelial markers, such as CK19. Recent studies using genetically engineered mouse models of PDAC have shown that in fact both ductal and acinar cells can give rise to PDAC precursor lesions by oncogenic KRAS activation.(4) Furthermore, transient acinar-to-ductal metaplasia was observed in mouse models, with
  • 10. 4 reversible phenotypic and molecular changes that persisted in the presence of chronic inflammation or oncogenic KRAS activation. Although there is also evidence for this phenomenon in resected human PDAC surgical specimens, it has been argued that the metaplastic lesions may be intraductal spread of pre-existing PDAC and/or its precursor lesions. Genetic Landscape of Pancreatic Cancer The genetic landscape of PDAC is characterized predominantly by mutations in four major driver genes, listed in the order of decreasing frequency: KRAS, CDKN2A, SMAD4, and TP53. Frequent alterations in these genes were first identified by candidate gene sequencing and have since been corroborated repeatedly by multiple large exome and genomic sequencing studies of PDAC.(5) Activating mutations of oncogene KRAS are seen in more than 90% of PDACs, and inactivating mutations of tumor suppressor genes, CDKN2A, SMAD4 and TP53 in 50-80% of PDACs.(2) An additional 32 recurrent ‘passenger’ mutations – defined as those co- occurring with driver mutations without conferring additional growth advantage – were also identified, including but not limited to ARID1A, RNF43, TGFBR1, TGFBR2, MLL3, MKK4, KDM6A, PREX2, RB1 and CCND1, at lower frequencies in approximately 10% of PDAC tumors, highlighting the significance of tumoral heterogeneity (Table 1).(2, 4) It will be important to fully characterize the functional significance of these passenger gene mutations as they represent genetic differences among PDACs that may be exploited clinically. Precursor Lesions At least three histologically distinct precursor lesions of PDAC have been described so far, consisting of pancreatic intraepithelial neoplasm (PanIN), and two types of mucinous cystic lesions including intraductal papillary mucinous neoplasm (IPMN) and mucinous cystic neoplasm (MCN). These precursor lesions are further characterized histologically and graded
  • 11. 5 according to their degree of dysplasia as lesions of low-grade versus high-grade dysplasia (Figure 1). Targeted sequencing of PanIN lesions along with their matched corresponding PDAC surgical resection specimens demonstrated that the same four driver genes are mutated in PanIN at very high frequencies as observed in PDAC. Comprehensive exome and whole genomic sequencing studies also confirmed these findings, establishing PanIN as the canonical precursor lesion of PDAC.(5) Similarly, shared mutations were also seen with mucinous cysts and their matched corresponding PDACs. Targeted sequencing of IPMNs identified shared mutations in genes GNAS and KRAS, and exome sequencing of IPMNs and MCNs identified shared mutations in RNF43, indicating that cystic neoplasms represent additional precursor lesions of PDAC that employ different progression pathways.(4) Remarkably, mutational analysis comparing PanINs of different grades revealed a positive correlation between the PanIN grade and the frequencies at which driver gene mutations are found.(5) Furthermore, it revealed a sequential pattern in which mutations found to accumulate in a predictive order following the PanIN grade. High-sensitivity methods to detect KRAS mutations showed their involvement in more than 99% of all PanIN-1 lesions, suggesting that oncogenic transformation of KRAS is most likely the initiating step in the development of pancreatic cancer.(6) While KRAS mutations are found across all grades of PanINs and invasive PDACs, the proportion of cells harboring the mutation increases with higher PanIN grade, indicating a clonal expansion of cells carrying the mutation.(6) In addition to oncogenic KRAS, inactivating mutations in CDKN2A can be seen in PanIN-2 and again at a higher frequency in PanIN-3.(5) Similarly, SMAD4 and TP53 mutations are additionally found in PanIN-3 and in
  • 12. 6 invasive PDACs, with both SMAD4 and TP53 mutations occurring at higher frequencies in invasive PDACs. These findings may be explained by a linear progression model of pancreatic cancer development, in which mutations are acquired in a gradual, step-wise pattern. By sequencing primary PDACs and their matched metastatic tumors, it was estimated that the linear progression from a nascent pancreatic cell acquiring an initiating driver gene mutation to the ultimate development of invasive PDAC would take 10 or more years.(7) This notion is consistent with the observation that nearly 33% of pancreata seen in autopsy series contain PanINs, suggesting that PanINs are quite common and generally do not progress to an invasive cancer.(6) In contrast, an alternative model termed chromothripsis proposes a punctuated evolution of pancreatic cancer, in which catastrophic genomic events involving structural alterations cause simultaneous inactivation of multiple driver genes in a single cell cycle. In support of this model, whole genome sequencing of primary tumors demonstrated two-thirds of PDACs having complex structural variations that, in a subset of cases, simultaneously inactivated multiple driver genes.(8) In the same study, many tumors did not harbor the predicted sequence of mutations, suggesting that these mutations may be acquired in a stochastic fashion consistent with a chromothripsis model. Still, a third model that combines both linear progression and punctuated evolution is entirely plausible. Distinguishing among these mechanistically distinct yet mutually non-exclusive models has clinical importance, since under a linear progression model which predicts a slow and gradual progression of disease, clinical efforts are best geared toward improving methods for screening and early detection of pancreatic cancer, whereas under a punctuated evolution model, an emphasis on enhancing systemic therapy is more appropriate.
  • 13. 7 Mutational Processes To fully understand the pathophysiology of PDAC, it is essential to delineate the mutational processes that are operative in the development of PDAC. Framing pancreatic cancer in familiar evolutionary terms can facilitate a mechanistic understanding of how mutations arise in the first place. In Darwinian evolution, mutations occur purely stochastically in dividing cells at an expected somatic mutation rate of three single nucleotide variants per cell division.(6) In the case of the pancreas which does not comprise of highly proliferative tissues, the probability of a pancreatic cell acquiring an initiating driver gene mutation by random chance alone is exceedingly low, and can be expected to largely depend on the total number of cell divisions performed over the lifetime of the dividing cell. Not surprisingly, statistical analysis of various types of human cancers, including PDAC, revealed a strong correlation between lifetime cancer risk and the number of cell divisions performed by adult stem cells of a given organ.(9) This finding lends support to the well-established finding that patient age is a major risk factor for the development of PDAC. Indeed, most pancreatic cancer patients are diagnosed at beyond age 50, with peak incidence occurring in the seventh and eighth decades of life.(2) However, the relative contribution of intrinsic factors (e.g. stochastic mistakes taking place during DNA replication) versus extrinsic factors (e.g. patient exposure to carcinogens or radiation) to lifetime risk remains a point of contentious debate. By whole-genome and RNA sequencing of resected PDAC surgical specimens, Connor et. al identified four distinct mutational processes acting on the PDAC genome.(4, 10) Those related to increasing age and number of cell divisions were the most prevalent, accounting for approximately 70% of all mutational signatures observed. To lesser degrees, mismatch repair (MMR) defects accounted for 2%, homologous recombination (HR) defects accounted for 11%,
  • 14. 8 and a process of unknown etiology termed “Signature 8” accounted for 15% of the mutational signatures. Tumors with MMR and HR defects characteristically showed biallelic inactivation of genes essential for the respective DNA repair processes, including MSH2, BRCA1, BRCA2 and PALB2. Also, one allele was often lost in the germline, which explains the involvement of the same genes in familial pancreatic cancers. Of note, tumors with MMR defects, owing to their microsatellite instability, exhibited higher burdens of somatic mutations and increased transcription of antitumor immune markers as determined by RNA sequencing, which may translate to a greater responsiveness to immunotherapy. Tumoral Heterogeneity The complex genetic landscape of PDAC is complicated by significant tumoral heterogeneity, which can be further categorized into intratumoral and intertumoral heterogeneity. Intratumoral heterogeneity, which describes genetic heterogeneity that exists among cells of a single tumor, is a well-recognized prognostic factor and an important cause of therapeutic resistance in pancreatic cancer. The concept of intratumoral heterogeneity first became apparent in lineage tracing studies of primary PDACs and matched metastatic tumors, which determined that metastatic tumors arise from distinct subclonal outgrowths from the primary lesion, all of which likely diverged from a single parental clone.(7) Intratumoral heterogeneity in a patient can manifest in three forms: [1] subclonal heterogeneity within a primary tumor, where a founder clone gives rise to various subclones by acquiring additional mutations, [2] subclonal heterogeneity within a metastasis, where a metastasis-initiating cell gives rise to its descendant subclones in a similar fashion, and [3] subclonal heterogeneity of metastasis-initiating cells within a primary tumor, where metastasis-initiating cells share common ancestors but possess distinct mutations that confer varying degrees of metastatic potential.(6)
  • 15. 9 Intertumoral heterogeneity describes genetic heterogeneity that exists among tumors of same histological type occurring in different patients, and it has been well-described in pancreatic cancer. To characterize the intertumoral differences systematically, several classification systems have been proposed based on genomic, transcriptomic, and immunohistochemical analyses. Molecular Subtyping of Pancreatic Cancer Waddell et al. classifies PDAC into four major subtypes based on patterns of structural variation identified from their genomic analysis.(11) In their study, 20% of tumors had ‘stable’ genomes with fewer than 50 structural variants, 36% of tumors had ‘scattered’ structural events with 50-200 variants, 14% of tumors had ‘unstable’ genomes with more than 200 structural variants suggestive of defects in DNA maintenance, and lastly, 30% of tumors had a ‘locally rearranged’ pattern with fewer than 50 structural variants localized to 1-3 chromosomes which typically result from amplifications that encompass oncogenes or genomic catastrophes such as in the case of chromothripsis. Interestingly, the ‘unstable’ subtype was predictive of platinum and poly (ADP-ribose) polymerase (PARP) inhibitor responsiveness. Transcriptomic studies of PDAC have also identified different molecular subtypes of PDAC with prognostic and therapeutic implications, resulting in a number of classification systems that differ based on the input material used and assumptions made for each study. Using microarray expression analysis of microdissected epithelium, Collison et al. classifies PDAC into three subtypes termed ‘classical’, ‘quasimesenchymal’ and ‘exocrine-like’.(12) Notably, the classical subtype was predictive of therapeutic response to erlotinib, while the quasimesenchymal subtype was negatively prognostic and predictive of therapeutic response to gemcitabine. In a similar study, Bailey et al. analyzed transcriptomic data from bulk tissue
  • 16. 10 containing the tumor microenvironment, and identified an additional ‘immunogenic’ subgroup based on presence of stromal immune cell populations.(13) Still, Moffitt et al. proposed a new classification system by excluding transcripts from presumed normal pancreas from their analysis, and identified two tumoral subtypes – ‘classical’ versus ‘basal-like’ – as well as two stromal subtypes – ‘normal’ versus ‘activated’.(14) Tumors corresponding to ‘basal-like’ subtype and ‘activated’ stromal subtype were independently and additively negatively prognostic. The basal type was also more responsive to chemotherapy on retrospective analysis. Although large-scale genomic and transcriptomic analyses have greatly elucidated the intertumoral heterogeneity of PDAC defining its molecular subtypes and established a foundation for developing precision medicine, applying this knowledge clinically has been limited by the common lack of access to complex tumor tissue biobanking and sequencing platforms for most clinicians. To this end, Noll et al. asked whether immunohistochemical (IHC) analysis, which is a far more accessible and technically feasible form of testing for clinicians at large, could be used to subtype pancreatic tumors by protein expression, and determined two IHC markers – HNF1A and KRT81 – for the differentiation of Collison subtypes.(15) Specifically, HNF1A-positive tumors correlated to the exocrine-like subtype, KRT81-positive tumors to quasimesenchymal subtype, and IHC-negative tumors to classical subtype. In addition, their study identified CYP3A expression as a novel mechanism of drug resistance, found at higher levels in exocrine-like tumors but inducible in all subtypes. In 2009, Farrell et al. reported the predictive value of an IHC-based assay for guiding precision medicine treatment of pancreatic cancer. In a phase III adjuvant therapy trial of 538 patients with early pancreatic cancer, the expression of human equilibrative nucleoside transporter (hENT1) – a key mediator of cellular uptake of gemcitabine – measured by IHC
  • 17. 11 analysis of tumor microarrays was associated with increased overall survival and disease-free survival in patients who received gemcitabine, but not in those who received 5-FU, demonstrating hENT1 as a predictive biomarker for gemcitabine efficacy in patients with early pancreatic cancer.(16) Deranged Signaling Pathways / Molecular Aberrations The full mutational landscape of pancreatic cancer is highly complex and diverse. PDACs contain an average of 63 genetic alterations, the majority of which consists of infrequent mutations found in fewer than 10% of PDACs.(2, 17) Nonetheless, many cases of these low- frequency targets appear to be alternative perturbations of the same core signaling pathways that are commonly deranged across all PDAC subtypes. By interrogating the exome of 24 PDACs, Jones et al. determined 12 core signaling pathways consistent with the hallmarks of cancer previously described by Hanahan and Weinberg, although the specific genes and the number of genes altered in each pathway differed from patient to patient.(17, 18) Included among the pathways were those affected by well-known driver genes, such as TP53 in DNA damage response and SMAD4 in TGFβ signaling. Some pathways, such as RAS-ERK signaling and DNA damage response, were predominated by a single frequently mutated gene, while others, such as integrin signaling, regulation of invasion, homophilic cell adhesion and GTPase- dependent signaling, involved many different genes. Biankin et al. further enriched our knowledge of commonly deranged pathways by next-generation exome sequencing, shedding light on the deregulation of axon guidance (SLIT and ROBO2), DNA damage repair (ATM) and chromatin modification (EPC1) in PDAC, which were formerly unappreciated.(19) Aberrant autocrine and paracrine signaling cascades ultimately promote pancreatic cancer cell proliferation, migration, invasion, and metastasis.(2) Numerous cytokines, such as
  • 18. 12 transforming growth factor-α (TGFα), insulin-like growth factor 1 (IGF1), fibroblast growth factors (FGFs) and hepatocyte growth factor (HGF), and their respective tyrosine kinase receptors, lead to pathologic activation of multiple pathways that confer pancreatic cancer cell mitogenic self-sufficiency. These signaling cascades also act to promote cancer cell migration and invasion of both local and distant sites, leading to metastasis. Pancreatic cancer cell proliferation is further enhanced by pathologic activation of anti-apoptotic and pro-survival pathways, such as signal transducer and activator of transcription 3 (STAT3), nuclear factor-κB (NF-κB) and AKT. Reactivation of genes involved in early development, such as WNT, SHH and NOTCH, can also be seen in a subset of PDAC. Pathway derangements in PDAC are numerous, and deconstructing their downstream effects is further complicated by significant crosstalk between pathways creating synergistic outcomes.(6) p53 normally cooperates with receptor SMADs to activate TGFβ-induced transcription by forming complexes that bind separate cis-enhancer elements on a target gene promoter. In PDAC, oncogenic KRAS interferes with TGFβ signaling by degrading SMAD4 and inhibiting p53 by blockade of its amino-terminal phosphorylation. Furthermore, oncogenic KRAS and mutant p53 form pathologic complexes that in turn inhibit p63, which normally acts to oppose TGFβ-dependent cell migration, invasion and metastasis. Collectively, these findings indicate that deranged pathways in pancreatic cancer exist not as independent processes but rather as a complex tumorigenic network altering the systems biology of the cell.(6) Tumor Microenvironment A hallmark of PDAC is its abundant and dense collagenous stroma, which may account for up to 90% of the total tumor volume. The tumor microenvironment of PDAC consists of a highly complex assembly of diverse cell types, including pancreatic stellate cells (PSCs),
  • 19. 13 immune cells, endothelial cells and nerve fibers, which are influenced by the extracellular matrix composed of matricellular proteins, fibrillar collagen, fibronectin, hyaluronic acid and a wide range of cytokines, such as TGFβ, FGF, epidermal growth factor (EGF) receptor ligand, vascular endothelial growth factor (VEGF) and connective tissue growth factor (CTGF). There is now abundant evidence for the prominent role of pancreatic cancer-associated stroma in tumor progression by actively promoting tumor growth, invasion and metastasis. Recently, a protective effect of some of the stromal components contributing to a physical containment of cancer cells has also been suggested. The dual function of PDAC stroma as both a tumor promoter and a suppressor suggests that its pathogenic role may arise from a loss of balance between epithelial cells and stroma. While normal extracellular matrix has the capacity to restrain tumor growth through the histone demethylase JMJD1a, desmoplastic stroma consists of aberrant matrix that is stiff with thickened collagen fibers and expresses p-MLC2 that contributes to tumor progression.(20, 21) PSCs are major drivers of the desmoplastic reaction in PDAC, wherein pancreatic tissue injury leads to PSC activation and trans-differentiation into α- smooth muscle actin expressing myofibroblast-like cells secreting collagen-type I, matrix metalloproteinases (MMPs) and tissue inhibitor of metalloproteinases (TIMPs) that remodel the extracellular matrix. PSC activation can be triggered by various cytokines and stimuli, including platelet-derived growth factor (PDGF), TGFβ1, FGF, EGF, tumor necrosis factor-α (TNFα), interleukin-1 (IL-1), ethanol, endotoxins, hypoxia, pressure and oxidative stress, many of which are produced by pancreatic cancer cells, endothelial and immune cells of the microenvironment. Once established, PSC activation is maintained in an autocrine fashion. The resulting fibrous stroma is a severely hypoxic, nutrient-deprived environment that promotes tumor aggressiveness by activation of hypoxia-inducible factor-1a. In addition, activated PSCs directly promote
  • 20. 14 proliferation of cancer cells by secreting mitogenic factors such as stromal-derived factor-1, PDGF, EGF, IGF-1 and FGF which activate MAPK- and AKT-signaling cascades.(22) Another key feature of the PDAC microenvironment is its highly immunosuppressive composition. Once the tumor is established, the tumor microenvironment is immunosuppressed by several mechanisms, including an accumulation of regulatory T cells, M2 type tumor- associated macrophages and myeloid-derived suppressor cells (MDSCs). Activated KRAS in tumor cells directs the transcription of granulocyte-macrophage colony-stimulating factor (GM- CSF), an inflammatory cytokine that promotes recruitment and trans-differentiation of myeloid progenitor cells into MDSCs which in turn suppress the immune surveillance function of CD8+ T cells.(23) Tumor cells also stimulate the expression of IP-10 (CXCL10) in PSCs which attract CXCR3+ regulatory T cells to the tumor milieu.(24) PSCs also secrete CXCL12 which attracts CD8+ T cells away from the juxtatumoral stromal compartment, reducing their chance to interact with cancer cells.(25) In addition, various cell types within the tumor microenvironment secrete numerous cytokines that support the immunosuppressive phenotype, including IL-1b, IL-4, IL-5, IL-6, IL-8, IL-10, IL-13, TNFα, TGFβ, FGF, PDGF, MMPs, thymic stromal lymphopoietin (TSLP), interferon-γ (IFN-γ) and VEGF.(23) Ultimately, the PDAC microenvironment appears to constitute a biological space of immune privilege where cancer cells are protected from immune surveillance, as opposed to rendering T cells dysfunctional as mechanisms to bypass mechanisms of T cell suppression can promote intratumoral infiltration of cytotoxic T cells and uncover latent immune responses.(26, 27) Further research on the dynamic intersection of pancreatic cancer and its tumor microenvironment is of great clinical importance as it will likely provide answers to improving delivery of chemotherapy and developing effective immunotherapy.
  • 21. 15 Metabolic Reprogramming Successful pancreatic cancer cell survival and proliferation depends on its ability adapt to a severely hypoxic and nutrient-deprived tumor microenvironment. Indeed, pancreatic cancer cells are known to employ various metabolic changes through mechanisms that are mainly driven by the expression of oncogenic KRAS and hypoxia-inducible transcription factor 1α (HIF1α).(2) Oncogenic KRAS induces overexpression of glucose transporter 1, hexokinase 1 and hexokinase 2, which significantly increases glucose uptake by pancreatic cancer cells. The increased levels of glucose are funneled through aerobic glycolysis to provide substrates for ATP production such as pyruvate as well as for the synthesis of nucleic acids, proteins, and fatty acids. This process in PDAC is uncoupled from the tricyclic acid (TCA) cycle and electron transport chain (ETC) via HIF1α-mediated induction of pyruvate dehydrogenase kinase 1, which phosphorylates and inactivates pyruvate dehydrogenase, thereby limiting the conversion of pyruvate to acetyl-CoA needed for the TCA cycle. The uncoupling of events results in increased production of lactate, which in turn becomes an important nutrient for less hypoxic cancer cells, and reduces the production of reactive oxygen species by ETC. Moreover, oncogenic KRAS promotes macropinocytosis in cancer cells as a major mechanism for the uptake of extracellular proteins to meet cellular requirements for glutamine and other amino acids. Similarly, HIF1α activates the autophagy-lysosome system, a self-degrative process for cytoplasmic components including organelles and macromolecules, to maintain intracellular energy supplies. In xenograft mouse models of PDAC, pharmacologic inhibition of these processes substantially delayed tumor growth.
  • 22. 16 Immune Response in Pancreatic Cancer is Unclear Development of immunotherapies has revolutionized the treatment options for many types of cancers, including but not limited to melanoma, renal and lung cancers. These therapies rely on potentiating pre-existing tumor-specific T cells by blockade of immune checkpoints, which are inhibitory pathways in place to maintain self-tolerance and modulate physiological immune responses to minimize collateral tissue injury. The same pathways are exploited by tumors to gain immune resistance against tumor-specific T cells. Some cancers, notably PDAC, are refractory to immunotherapies, and it remains unclear why. The failure of numerous immune checkpoint inhibitors to advance through clinical trials for treatment of PDAC created a preconceived notion in the scientific community that PDACs are poorly immunogenic tumors. However, an increasing number of studies have now shown prominent T cell infiltrates in the vast majority of biopsies from PDAC patients and identified unique neoantigen qualities in long- term survivors, indicating that a meaningful immune response in PDAC is achievable.(28, 29) However, research in this area has been hampered by the lack of pre-clinical physiologic models of PDAC that are suited to study anti-tumor immune response. Pre-clinical Modeling of Pancreatic Cancer “KP-C” (KrasLox-STOP-Lox-G12D ; P53Lox-STOP-Lox-R172H/+ ; Pdx1-Cre) mice have been widely used to investigate pancreatic cancer biology. Although this model has been greatly informative regarding the genetic landscape of PDAC, it is ill-suited for the study of cancer immunology on two levels. First, tumors develop aggressively in these mice, rapidly progressing to fatal metastatic disease predominantly by 6 weeks of life. This creates a practical challenge in investigating early disease when meaningful tumor-immune cell interactions may occur before significant stromal development and/or the onset of other mechanisms of immune
  • 23. 17 suppression.(30) Secondly, pancreatic tumors that develop in these mice are poorly antigenic, lacking neoantigen peptides which are critical for mounting anti-tumor T cell responses. In fact, depletion of T cells in KP-C mice using anti-CD4 and anti-CD8 antibodies had no effect on the progression of murine PDAC nor on the overall survival of these mice.(31) Thus, most pancreatic cancer immunology studies have focused on murine and human PDAC cell lines, which have their own limitations.(32) Namely, monolayer cell lines lack the structural sophistication and functional differentiation of cells seen in vivo, and cannot recapitulate the tumor microenvironment in mouse xenograft studies. Cell line-derived three-dimensional spheroid cultures attempt to address this issue, but are difficult to propagate in spheroid form, limiting longitudinal investigations. Furthermore, none of the cell-line derived models support the growth of untransformed, non-neoplastic cells. Instead, they inevitably become monoclonal over time by in vitro selection of the most aggressive clones, resulting in a loss of genetic heterogeneity seen in primary tumors. Patient-derived xenograft (PDX) mouse models, which are established by implanting a piece of surgically resected tissue from a patient under the dermis of immunocompromised mouse hosts, are inherently more physiologic but are cost-prohibitive and excessively time-consuming, commonly taking upwards of 6 months to generate sufficient sizes of mouse colonies, which is outside clinically meaningful timeframes for any approach to personalized medicine for most pancreatic cancer patients. A recent breakthrough in translational pancreatic cancer research has been the development of organoid models of pancreas using human and mouse pancreatic tissues for pre- clinical modeling of PDAC. Organoids, comprising of complex clusters of multiple cell types derived from the tissue of interest, can recapitulate the intricate spatial architecture of the progenitor organ structure and perform functions of the organ such as secretion or contraction.
  • 24. 18 Since a robust method for production of self-renewing intestinal organoids was first reported in 2009, tumor organoid models have been widely adopted for multiple organ systems.(33) In 2015, Boj et al. recently described methods for reliably generating human and mouse PDAC organoids using surgical resection specimens as well as endoscopic ultrasound-guided fine needle biopsy (EUS-FNB) specimens.(34) PDAC organoids derived in this manner could recapitulate the natural history of human PDAC when orthotopically transplanted into immunocompromised mice, forming early PanIN-like lesions that progressed to invasive pancreatic cancer with robust stromal response. The ability to generate organoid cultures from FNB specimens is a major advantage, since it enables investigators to capture the full spectrum of PDAC ranging from early premalignant lesions to late metastatic cancers, as opposed to surgical resection specimens which account for fewer than 20% of patients diagnosed with PDAC who are surgical candidates. The organoid model is physiologic yet possesses all the desirable intrinsic properties of an in vitro system. PDAC organoid cultures can be propagated in vitro for expansion of starting material, which is often the limiting factor for tissue-consuming studies such as deep sequencing, and cryopreserved indefinitely without losing genetic heterogeneity. They are highly tractable, amenable to genetic manipulation and high-throughput assays. Moreover, in contrast to PDX mouse models, organoid cultures can be established rapidly in sufficient quantities for studies in just 2-4 weeks from the time point of acquiring patient tissues, permitting a personalized approach to pancreatic cancer medicine to investigate patient-specific tumor biology, evaluate prognosis and guide therapy in real time. II. STATEMENT OF PURPOSE In this study, we sought to develop murine and human organoid models of PDAC to investigate the biology of pancreatic cancer immune response. Our aims were mainly two-fold:
  • 25. 19 1. Development of an immunogenic murine PDAC organoid model to study antigen- specific anti-tumor T cell responses in both in vivo and in vitro setting. 2. Creation of a clinically annotated library of validated, patient-derived PDAC organoid lines as tools for studying human pancreatic cancer immunology. III. METHODS Acquisition of human specimens Human pancreatic cancer tissues were obtained from patients undergoing endoscopic ultrasound-guided fine needle biopsy (EUS-FNB) or surgical resection at Yale New Haven Hospital. Some of the surgical resection specimens were used to create patient-derived xenograft (PDX) mouse models, which subsequently became available as a secondary source of patient- derived tissues for generation of organoids. Tissues were determined to be tumoral or normal by evaluation of on-site clinical pathologist. Written informed consent was obtained from all patients prior to tissue acquisition. This study was reviewed and approved by the Institutional Review Board of Yale University. All EUS-FNB specimens were provided by James Farrell who also performed the biopsies. All surgical resection specimens were histologically evaluated and provided by Marie Robert. PDX mouse models of PDAC were previously established by Ryan Sowell in Kaech laboratory. Isolation and culture of murine pancreatic organoids Murine pancreatic organoids were generated using normal or pre-neoplastic pancreatic tissues from C57BL/6 mouse and KP-NINJA (KrasLox-STOP-Lox-G12D ; P53flox/flox ; inversion induced joined neoantigen) mouse, respectively. Detailed procedures for isolation and propagation of murine pancreatic organoids were adapted from Boj et al., 2015 and Huch et al., 2016. Briefly, mouse pancreas was dissected and minced into sub-millimeter pieces before enzymatic digestion
  • 26. 20 with collagenase XI (0.125 mg/mL, Sigma-Aldrich), dispase II (0.125 mg/mL, Thermo Scientific) and DNase I (0.1 mg/mL, Sigma-Aldrich) in advanced DMEM/F12 medium (Life Technologies) supplemented with FBS (2.5%), Glutamax (1X, Thermo Scientific) and Antibiotic-Antimycotic (1X, Thermo Scientific) for 1-3 hours at 37ºC in a tissue dissociator until visual confirmation of pancreatic ducts which were manually picked for ductal enrichment under a dissecting microscope. Harvested ductal fragments were embedded in growth factor reduced Matrigel (Corning) and cultured in complete murine organoid growth medium, consisting of advanced DMEM/F12 supplemented with Glutamax (1X), HEPES (10 mM, Life Technologies) and Antibiotic-Antimycotic (1X), Rspo1-conditioned medium (10% v/v), human noggin (0.1 µg/mL, Peprotech), B27 supplement minus vitamin A (1X, Thermo Scientific), N-acetyl cysteine (1.25 mM, Sigma-Aldrich), nicotinamide (10 mM, Sigma-Aldrich), human gastrin I (10 nM, Sigma-Aldrich), mouse EGF (50 ng/mL, Thermo Scientific), human FGF-10 (100 ng/mL, Peprotech) and A83-01 (500 nM, Tocris Bioscience). Y-27632 (10.5 µM, Tocris Bioscience) was added for initial organoid cultures following isolation from primary tissue, single cell dissociation, or thawing from cryopreservation. Murine organoid models were characterized by in vivo transfer for tumor formation in C57BL/6 mouse and immunohistochemical analysis of resulting tumors. KP-NINJA mouse model was previously established by Nikhil Joshi. All procedures outlined above were performed by the author. Isolation and culture of human PDAC organoids Human PDAC organoids were generated using patient-derived tissues from EUS-FNB, surgical resection, or pre-established PDX mouse models. Detailed procedures for isolation and propagation of human PDAC organoids were adapted from Boj et al., 2015 and Huch et al., 2016. Briefly, tissues were minced into sub-millimeter pieces before enzymatic digestion with
  • 27. 21 collagenase II (5 mg/mL, Thermo Scientific), dispase II (0.125 mg/mL) and DNase I (0.1 mg/mL) in advanced DMEM/F12 medium supplemented with FBS (2.5%), Glutamax (1X) and Antibiotic-Antimycotic (1X) for 1-3 hours at 37ºC in a tissue dissociator until tissues become submacroscopic. Cells are embedded in growth factor reduced Matrigel and cultured in complete human organoid growth medium, consisting of advanced DMEM/F12 supplemented with Glutamax (1X), HEPES (10 mM) and Antibiotic-Antimycotic (1X), Wnt3a-conditioned medium (50% v/v), Rspo1-conditioned medium (10% v/v), human noggin (0.1 µg/mL), N2 supplement (1X, Thermo Scientific), B27 supplement minus vitamin A (1X), N-acetyl cysteine (1.25 mM), nicotinamide (10 mM), human gastrin I (10 nM), human EGF (50 ng/mL, Peprotech), human FGF-10 (100 ng/mL, Peprotech) and A83-01 (500 nM). Y-27632 (10.5 µM) was additionally added for initial organoid cultures following isolation from primary tissue, single cell dissociation, or thawing from cryopreservation. Human organoid models were characterized by Sanger sequencing of KRAS and P53, in vivo transfer for tumor formation in immunodeficient NOD scid gamma (NSG) mouse (The Jackson Laboratory) and immunohistochemical analysis of resulting tumors. FNB-derived organoid lines were established by the author. Surgical resection- derived and PDX mouse-derived organoids were established by collaborative effort of Prashanth Gokare and the author. Characterization of patient-derived organoids was performed by collaborative effort of Prashanth Gokare and the author. Isolation of primary murine PDAC cell lines Primary murine PDAC cell lines were isolated from pancreatic tumors harvested from KP-C (KrasLox-STOP-Lox-G12D ; P53flox/flox ; Pdx1-Cre) mouse. Resected tumors were minced into sub-millimeter pieces before enzymatic digestion with trypsin-EDTA (0.25%, Life Technologies) and collagenase IV (1 mg/mL, Worthington Biochemical) in HBSS buffer (1X,
  • 28. 22 Life Technologies) for 30 min at 37ºC in a tissue dissociator, after which the digestion reaction was quenched using cold FBS. Cells were passed through a 40 µm filter to prepare single cell suspension and washed twice prior to cell culture in complete DMEM. All procedures outlined above were performed by Brittany Fitzgerald. Genetic manipulation of murine pancreatic organoids Murine pre-neoplastic pancreatic organoids isolated from KP-NINJA mouse model were in vitro transformed into neoplastic organoids using LV-rtTA-Cre-iRFP670. Detailed procedures for genetic manipulation of organoids were adapted from Huch et al., 2016. Briefly, a single cell suspension of organoids was prepared by pooling 3 confluent wells of a 24-well plate, removal of Matrigel, and digestion of organoids in TrypLE Express (1X, Life Technologies) and DNAse I (0.1 mg/mL) for 5 min at 37ºC with vigorous pipetting every 2 min. After washing, cells were resuspended with concentrated lentivirus and spinoculated at 600 G for 1 hour at 32ºC, followed by incubation for 6 hours at 37ºC. iRFP670 labeling of infected organoid fragments could be visualized by fluorescence microscopy 2-3 days after infection. After expansion, organoids were analyzed by flow cytometry for expression of iRFP670 and sorted for the brightest 10% of cells expressing iRFP670. For the expression of programmed, GFP-tagged neoantigens, a subset of transformed organoids was also infected with Ad-FLPo by spinoculation followed by incubation as described above. Ad-Cre was used as a negative control. Alternatively, a subset of organoids was treated with doxycycline and tamoxifen in vitro to achieve the same effect. After expansion, organoids were analyzed by flow cytometry for expression of GFP and sorted for the brightest 10% of cells expressing GFP. Lentiviral and adenoviral transformations of organoids were performed by the author. Creation of immunogenic organoid lines by treatment with doxycycline
  • 29. 23 and tamoxifen was performed by Gena Foster. Flow cytometry and cell sorting of transformed organoids were performed by collaborative effort of Gena Foster and the author. In vivo mouse assays Murine and human pancreatic organoids were characterized by subcutaneous injection of organoids for in vivo tumor formation in C57BL/6 mouse or NSG mouse, respectively. To standardize injections, organoids were first dissociated into single cells and seeded at concentration of 2.5 x104 cells per well in 24-well plate format. Organoids were then expanded to 80-90% confluency in a period of 5-10 days depending on the organoid line. For each mouse injection, organoids were pooled from 6 confluent wells for a total of approximately 5 x 105 cells, broken down into organoid fragments by vigorous pipetting using 200 µL pipette tips, and finally resuspended in 50 µL of Matrigel diluted 1:1 with cold PBS. Mice were anesthetized using isoflurane for injections and subsequently monitored for subcutaneous growth of tumors by caliper measurement every 2 days. Mice were euthanized promptly when tumors reached 1 cm in size or whenever a humane concern developed. Resulting tumors were harvested and analyzed by immunohistochemistry. All procedures outlined above were performed by the author. To test the effects of programmed neoantigens on tumor growth in vivo, neoantigen- negative, neoantigen-positive and weakly neoantigen-positive murine PDAC organoids were injected subcutaneously into C57BL/6 mice as described above. For creation of a weakly neoantigen-positive organoid line with only 10% of its cells expressing neoantigens, neoantigen- positive organoids were diluted 1:9 with neoantigen-negative organoids. A cohort of mice also received retroorbital injections of luciferase-positive P14 splenocytes 24 hours prior to receiving organoid injection for co-transfer of antigen-specific T cells. Splenocytes were harvested from luciferase+ P14 mouse strain by homogenizing dissected spleen through a 70 µm strainer and
  • 30. 24 passing through a 27-gauge needle for single cell dissociation. Following centrifugation, RBC lysis was performed by incubation of cells in ACK Lysing Buffer (1X, Thermo Scientific) for 3- 5 min at room temperature. Cells were washed twice using RPMI medium prior to flow cytometry analysis for confirmation of tetramer-positive P14 CD8+ T cell population. For each mouse injection, 1 x 106 splenocytes were finally resuspended in PBS. Mice were anesthetized using isoflurane for each injection and subsequently monitored for subcutaneous growth of tumors by caliper measurement every 2 days. Mice receiving co-transfer of organoids and splenocytes were additionally monitored by IVIS Spectrum In Vivo Imaging System (PerkinElmer) for luciferase detection 24 hours after organoid injection and every 3 days thereafter. Mice were euthanized when tumors reached 1 cm in size or whenever a humane concern developed. Resulting tumors were harvested and analyzed by immunohistochemistry. Creation of immunogenic murine organoid lines and immunohistochemical analyses of resulting tumors were performed by collaborative effort of Gena Foster and the author. Isolation, flow cytometry and co-transfer of splenocytes were performed by collaborative effort of Brittany Fitzgerald and the author. Immunohistochemical analysis of tumors Primary and organoid-derived pancreatic tumors were analyzed by immunohistochemistry. Tissues were fixed in 10% formalin and embedded in paraffin. Sections were subject to H&E, RFP (600-401-379, Rockland) 1:1000, E-Cadherin (610182, BD Bioscience) 1:500, CK19 (Troma III, developed by Rolf Kemler, Max-Planck Institute of Immunobiology, Freiberg, Germany, and obtained from the Hybridoma Bank at the University of Iowa, Iowa City, Iowa, USA) 1:1000, Sox9 (AB5535, EMD Millipore) 1:1000, Muc5AC (ab212636, Abcam) 1:400, Phospho-Erk (4370S, Cell Signaling Technology) 1:400, and
  • 31. 25 Phospho-Mek (2338S, Cell Signaling Technology) 1:50. Immunohistochemical staining and imaging of tumor sections were performed by collaborative effort of Gena Foster and the author. Sanger sequencing of organoids Patient-derived organoids were sequenced for characteristic mutations in genes KRAS and P53 by Sanger sequencing as part of validation pipeline. Genomic DNA was prepared from organoids using DNeasy Blood & Tissue Kit (QIAGEN) and quantified using Nanodrop spectrophotometer. Regions of gene that are most frequently mutated were PCR amplified and sequenced using the same set of primers. Mutations at codons 12 and 13 of KRAS were determined by using sense primer: 5’AAAGGTACTGGTGGAGTATTTGATAG and antisense primer: 5’ACAAGATTTACCTCTATTGTTGGATC. Mutations at codon 61 of KRAS were determined by sense primer: 5’GGAAGCAAGTAGTAATTGATGGAGA and antisense primer: 5’GCATGGCATTAGCAAAGACTCA. Mutations in exon 5 of P53 were determined using sense primer: CAAGCAGTCACAGCACATGA and antisense primer: AACCAGCCCTGTCGTCTCT. Mutations in exon 6 of P53 were determined using sense primer: CAGGCCTCTGATTCCTCACT and antisense primer: AGACCTCAGGCGGCTCATAG. Mutations in exon 7 of P53 were determined using sense primer: ATCTCCTAGGTTGGCTCTGA and antisense primer: TGGCAAGTGGCTCCTGACCT. Mutations in exon 8 of P53 were determined using sense primer: CTCTTTTCCTATCCTGAGTA and antisense primer: CTGCTTGCTTACCCTGCTTA. PCR products were purified using QIAquick PCR Purification Kit (QIAGEN) and analyzed by gel electrophoresis for correct band size prior to sequencing. All procedures outlined above were performed by collaborative effort of Prashanth Gokare and the author.
  • 32. 26 Development of organoid-T cell co-culture model systems P14 CD8+ T cells were pre-activated and expanded out from harvested P14 mouse splenocytes by incubation of cells with GP33 peptide (0.1 nM, Anaspec) for 1 hour followed by incubation with human IL-2 (10 ng/mL, Peprotech) for 72 hours at 37ºC in complete RPMI medium supplemented with FBS (10%), HEPES (1X), non-essential amino acids (1X, Life Technologies), sodium pyruvate (1X, Life Technologies), 2-mercaptoethanol (55 µM, Sigma- Aldrich), penicillin-streptomycin (1X, Life Technologies) and Glutamax (1X), followed by cytometry confirmation and cell sorting of tetramer-positive P14 CD8+ T cells. Prior to co- culture, murine PDAC organoids were labeled with Calcein blue, AM (1X, Anaspec) while P14 CD8+ T cells were doubly labeled with Calcein blue, AM as well as Calcein green, AM (1X, Invitrogen). Neoantigen-positive and neoantigen-negative organoids were seeded in 10/20/30/40/50 uL volumes of Matrigel in 24-well plate format and cultured to 30% confluency. Prepared T cells were resuspended in complete RPMI medium at a concentration of 1 x 105 cells per 500 uL per well and were added carefully on top of Matrigel plugs after the removal of organoid growth medium. Co-cultures were subsequently monitored by live chambered fluorescence imaging on EVOS Cell Imaging System (Invitrogen) for 24 hours. All procedures outlined above were performed by collaborative effort of Prashanth Gokare and the author. IV. RESULTS KP-NINJA mouse model provides substrate for creation of immunogenic murine organoid models of PDAC To overcome the paucity of neoantigen peptides on pancreatic tumors that develop in standard KP-C mouse model and create an immunogenic murine PDAC organoid model, we generated pancreatic organoids from “KP-NINJA” (KrasLox-STOP-Lox-G12D ; P53flox/flox ; inversion
  • 33. 27 induced joined neoantigen) mouse model that has been genetically engineered to express glycoproteins GP33-41 and GP61-80 derived from lymphocytic choriomeningitis virus (LCMV) as CD8+ and CD4+ T cell neoantigens, respectively (Figure 2A). The neoantigens are tagged to the C-terminus of green fluorescent protein (GFP) functioning as a reporter for neoantigen expression. In order to ensure tight regulation of its expression, multi-layered genetic and drug- inducible mechanisms were engineered. This is critical as leaky expression of neoantigens during early developmental phase of mouse immune system can result in immune tolerance and loss of immunogenicity. The neoantigen cassette is inverted and flanked by non-compatible flippase recognition target (FRT) sites, requiring the action of flippase (FLP) recombinase to be properly expressed. The expression of FLP recombinase is regulated by a tetracycline response element (TRE), which requires reverse tetracycline-controlled transactivator (rtTA) – which can be introduced by any tissue specific promoter – plus doxycycline to be transcribed. The entire TRE- FLP recombinase cassette is floxed, requiring Cre recombinase mediated inversion to become poised for transcription. Introduction of Cre recombinase will also recombine KRAS and P53 resulting in the activation of oncogenic KRAS and deletion of P53 to drive tumorigenesis. Finally, FLP recombinase is fused to a mutated ligand binding domain of the human estrogen receptor (ER), requiring tamoxiphen to become stabilized and effective in the nucleus. As a result, KP-NINJA mouse model enables genetically and pharmacologically inducible expression of known neoantigens with precise temporal and spatial control. To generate normal pancreatic organoids from KP-NINJA mouse model, we adapted methods previously described by Boj et al.(34) Briefly, mouse pancreas dissection is followed by mechanical and enzymatic digestion to release ductal fragments, which are manually picked under a dissecting microscope for ductal enrichment (Figure 2B). The enriched ductal fragments
  • 34. 28 are then washed and seeded in basement membrane matrix (Matrigel) in 24-well plate format. Liquid medium containing essential components for pancreatic organoid culture is then added on top of congealed Matrigel plugs. After 5-10 days of tissue culture, budding of ductal fragments into spherical organoids can be observed. In vitro transformed murine pancreatic organoids form tumors that are histologically similar to early lesions found in human PDAC For in vitro transformation of KP-NINJA mouse-derived normal pancreatic organoids, a lentiviral construct encoding rtTA-Cre-iRFP670 was used for Cre recombinase-mediated activation of KRAS oncogene and deletion of P53. iRFP670 was included in the lentiviral construct as a fluorescent reporter for the expression of Cre recombinase, thereby labeling any transformed cell. Following lentiviral transformation, organoids were analyzed by fluorescence imaging and flow cytometry for expression of iRFP670 and sorted for the brightest 10% of cells expressing red fluorescent protein (RFP) (Figure 3A-B). Expression of GFP was included in the flow cytometry analysis to examine the possibility of undesirable leakiness of neoantigen expression, which did not occur in organoids. When lentivirus-transformed organoids versus untransformed normal pancreatic organoids were injected subcutaneously into the opposite flanks of same mouse, only the transformed organoids formed a tumor (Figure 3C). On histology, these tumors had numerous infiltrating well-differentiated ductal structures with epithelial lesions consisting of tall columnar cells with mucinous cytoplasm that are reminiscent of early lesions seen in human PDAC, as well as a robust stromal response with extensive fibrosis (Figure 3D). The ductal structures stained positively for RFP confirming transformed organoids as their cell of origin (Figure 4C). The stromal compartment did not stain for RFP, indicating that the stromal response is host-
  • 35. 29 derived. The ductal structures also stained positively for markers of epithelial and pancreaticobiliary origin, including E-cadherin, CK19 and Sox9, as well as for PDAC-associated tumor markers, such as Muc5AC and phosphorylated Erk and phosphorylated Mek which are downstream targets of oncogene KRAS in the MAPK/ERK signaling pathway. The histology of the organoid-derived tumors was comparable to that of primary murine pancreatic tumors that spontaneously form in KP-C mice (Figure 3D-E). Murine PDAC cell lines were generated from the primary pancreatic tumors of KP-C mice for comparative analysis. Tumors derived from injection of PDAC cell lines showed markedly different histology to that of tumors derived from injection of transformed organoids, notable for the absence of organized ductal structures in cell line-derived tumors, consistent with highly advanced, undifferentiated pathology as a result of a known caveat with monolayer cell lines, that is in vitro selection of aggressive clones (Figure 3F). Cell line-derived tumors also lacked a stromal response in contrast to organoid-derived tumors. Serial in vivo transfer of transformed murine pancreatic organoids results in progressively more advanced tumors We predicted that serial in vivo transfer of transformed organoids in mice by performing repeated rounds of organoid generation from tumors derived from organoid injections in a sequential fashion would lead to progressively more advanced tumors. After two rounds of in vivo transfer, the organoid-derived tumors showed increased features of high-grade dysplasia, including enlarged, hyperchromatic nuclei with prominent nucleoli, nuclear crowding and cell stacking (Figure 4A-B). Fluorescence imaging of organoids generated from tumors after one round of in vivo transfer not only confirmed retention of RFP label but also showed a more uniform labeling of organoids indicating in vivo enrichment for Cre recombinase-transformed
  • 36. 30 cells (Figure 4D). After third round of in vivo transfer, the tumors contained noticeably fewer organized ductal structures and appeared poorly differentiated. Expression of neoantigens in murine PDAC organoids elicits effective immune response in mouse For creation of immunogenic murine PDAC organoid lines, an adenoviral construct encoding FLP recombinase was used to genetically induce expression of GFP-tagged neoantigens in transformed organoids (Figure 5A). After adenoviral introduction of FLP recombinase, the organoids were analyzed by flow cytometry to confirm GFP expression and were sorted for the brightest 10% of cells expressing GFP (Figure 5B). An adenoviral construct encoding Cre was used as a negative control. To test whether expression of neoantigens can impact tumor growth, neoantigen positive versus neoantigen negative transformed organoids were injected subcutaneously into 3 cohorts of mice, where first cohort received neoantigen negative organoids, second cohort received neoantigen positive organoids, and third cohort received neoantigen positive organoids as well as retroorbital injections of luciferase-positive P14 CD8+ T cells, which have been genetically engineered to express the T cell receptor (TCR) specific for GP33, 24 hours prior to organoid injections. In vivo imaging of mice at 24 hours after organoid injections demonstrated accumulation of luciferase-positive P14 CD8+ T cells at the site of organoid injection in the third cohort (Figure 5C). Mice were monitored for subcutaneous growth of tumors for up to 30 days. None of the mice that received neoantigen- positive organoids developed tumors, while all 5 out of 5 mice that received neoantigen-negative organoids developed tumors, indicating immune clearance of neoantigen-expressing PDAC organoids (Figure 5D). Interestingly, upon injection of a mixture of neoantigen-positive and
  • 37. 31 neoantigen-negative organoids at a ratio of 1:9, tumors were able to form but were heavily infiltrated with immune cells on histology (Figure 6). Expression of neoantigens in murine PDAC organoids promotes T cell infiltration in T cell- organoid co-culture model Next, we developed a novel three-dimensional co-culture system to study interactions of murine T cells and PDAC organoids in vitro. Splenocytes were harvested from congenic P14 mice that have been genetically engineered to express TCRs specific for GP33, and were treated with IL-2 and GP33 peptide for expansion and pre-activation of constituent P14 T cells. Splenocytes were then analyzed by flow cytometry and sorted to prepare a pure population of pre-activated P14 CD8+ T cells. Cell-permeant live-cell staining dyes were used to distinguish cells in co-culture. PDAC organoids were labeled with green-fluorescent calcein AM dye, while P14 CD8+ T cells were doubly labeled with green- and blue-fluorescent calcein AM dyes. P14 CD8+ T cells were then introduced into the liquid medium of either neoantigen-positive or neoantigen-negative PDAC organoid cultures which were maintained in Matrigel plugs of different sizes to vary the amount of liquid media-Matrigel interface. Co-cultures were subsequently monitored under live fluorescence imaging for 24 hours. Within the first hour, there was significant clustering of P14 CD8+ T cells at the boundaries of Matrigel plugs which appeared to be a physical barrier to T cell entry (Figure 7A). Nonetheless, small yet increasingly large fractions of T cells could be observed to penetrate the Matrigel plugs containing neoantigen-positive PDAC organoids, such that by 24 hours there was a clearly noticeable difference in the amount of T cell infiltration between neoantigen-positive versus neoantigen- negative PDAC organoid co-cultures (Figure 7B).
  • 38. 32 Assembly of human PDAC organoid library To aid investigations of human pancreatic cancer immunology, we sought to build a clinically-annotated library of experimentally validated, patient-derived PDAC organoid lines at Yale. We obtained patient samples primarily from EUS-FNB specimens but also a smaller number of samples from surgical resection specimens and pre-established PDX mouse models of PDAC (Figure 8A). From October 2017 to May 2018, we successfully generated 21 patient- derived organoid lines from 24 FNB specimens, including one liver metastasis, all of which were pathology confirmed as PDAC for an overall organoid isolation efficiency of 87.5% (21/24). Established organoid cultures were validated as tumor organoids as opposed to normal pancreatic contaminants by in vivo transfer of organoids for tumor formation in immunocompromised (NSG) mouse, immunohistochemical analysis of organoid-derived tumors for common markers of PDAC, and Sanger sequencing of organoids for KRAS and P53, the two most commonly mutated genes in PDAC (Figure 8B-E). Under these criteria, 7 out of 9 patient-derived organoid lines tested to date have been successfully validated. Histology of tumors derived from organoid injections in mice closely matched the histology of their corresponding patient-derived primary tissues, confirming that organoids truly recapitulate the pathology of their source material (Figure 8C). The degree of dysplasia seen in organoid-derived tumors also correlated with the severity of disease of corresponding patients at the time of biopsy in three case studies (Fig. 8D). Organoids derived from a patient who had borderline resectable disease (Bx120817) formed tumors of moderately differentiated histology in agreement with the primary clinical pathology findings. In comparison, organoids derived from the primary tumor of a patient who had metastatic disease in the lungs (Bx111417) formed less differentiated tumors consisting of numerous disorganized ductal structures with
  • 39. 33 characteristic loss of lumens. Organoids derived from a metastatic lesion in the liver of a patient who had liver metastasis (Bx102417) formed the most poorly differentiated tumors with little to zero resemblance of normal ductal structures. V. DISCUSSION Development of effective immunomodulating therapies in pancreatic cancer remains elusive, largely owing to the lack of effective physiologic PDAC model systems for the study of biology of immune response against tumor. To fulfill this critical need, we have developed transplantable, immunogenic murine organoid models of PDAC that enable investigations of antigen-specific, anti-tumor T cell responses. By genetic engineering of inducible mutations in KRAS and P53, we were able to recreate the earliest genetic events in pancreatic tumorigenesis in vitro, and then follow the progression of disease in vivo after transplantations of organoids in mouse. Tumors derived from organoids in this way are histologically similar to early lesions found in human PDAC, demonstrating well-differentiated ductal structures infiltrating an extensive and dense fibrous stroma. Ability to recapitulate the tumor microenvironment which is fundamental to PDAC pathophysiology is essential for studying how stromal components impact immune response. Furthermore, by serial in vivo transfer of transformed murine organoids, we were able to generate transplantable organoid models that can reliably recreate discrete stages of PDAC progression from early precursor lesions to advanced invasive cancer. Recreation of early disease is especially critical for capturing meaningful tumor-immune cell interactions as immunosuppression is an early event in PDAC.(30) By introducing known T cell neoantigens in murine PDAC organoids, we were able to elicit robust and effective immune responses against neoantigen-expressing PDAC organoids in mouse transplant studies. High level of neoantigen expression in 100% of cells comprising PDAC organoids resulted in complete immune rejection
  • 40. 34 of organoid-derived tumor growth in mouse, whereas a low level of neoantigen expression by dilution of neoantigen-positive cells with neoantigen-negative cells by a factor of 10 permitted tumor growth albeit with increased immune infiltration, suggesting that both quality and quantity of neoantigen affect immune response. The expression of high-quality neoantigens in sufficient quantity before the development of an immunosuppressive tumor microenvironment may have been key to the successful immune clearance of PDAC organoids in this set-up. To facilitate mechanistic studies of antigen-specific T cell responses against tumor, we have developed an in vitro three-dimensional co-culture system that recapitulates interactions of T cells and PDAC organoids in vitro, where increased T cell infiltration of Matrigel plugs containing neoantigen- expressing organoids was observed. This system can be readily applied to study tumor organoid interactions with other important cell types such as tumor-associated macrophages or cancer- associated fibroblasts. Ability to generate effective immune responses against tumor using organoid models of PDAC challenges the widely conceived notion that PDAC is an inherently immunologically cold disease. Precision medicine is a newly emerging medical model that accounts for the unique biology of each patient, allowing for the development of targeted therapeutics against specific molecular mechanisms at play and a personalized approach to disease management based on the individual patient’s tumor characteristics. Since the advent of next-generation sequencing, our knowledge of biomolecular and genetic aspects of pancreatic cancer has grown exponentially in recent years, revealing novel insights into how precision medicine may be actualized, including an appreciation for the remarkable heterogeneity of PDAC. Integrated analyses of different ‘omics’ data sets enabled the categorization of pancreatic tumors into distinct molecular subtypes carrying prognostic and predictive values, paving the way to individualized therapy by
  • 41. 35 stratifying patients based on their tumor subtype for specific therapies. However, personalized medicine in pancreatic cancer has been difficult achieve due to the short median survival of pancreatic cancer patients and long turnaround times of standard PDX models. Development of patient-derived organoid models of PDAC has been revolutionary in this regard, as organoids can be derived from patients rapidly and analyzed within clinically relevant timeframes. In some cases, pharmacotyping of patient-derived organoids to generate drug-sensitivity profiles could be completed in as little as 6 weeks.(35) Moreover, organoids can model the full clinical spectrum of PDAC as they can be generated using small amounts of tissue from FNB specimens, thus removing the barrier to sampling non-surgical patients who account for more than 80% all pancreatic cancer patients, as opposed to standard PDX models that require surgical tissues. In our study, we have established a clinically annotated library of 30+ patient-derived PDAC organoid lines using FNB and surgical specimens. Our efforts to validate each patient-derived organoid line by tumor formation in mouse, immunohistochemistry and sequencing have been promising. Collectively, our data demonstrate that pancreatic organoids are an ideal model for the study of pancreatic cancer immune response. Our ongoing work includes using CRISPR/Cas9- based lentiviral systems in PDAC organoids to test and define genes that impact anti-tumor T cell responses with or without addition of immunomodulating agents in both in vitro co-cultures and in vivo mouse studies. We are also using organoid models of PDAC to investigate the pathogenic role of renalase – a recently discovered cytoprotective secreted flavoprotein that is upregulated in chronic pancreatitis and PDAC – and evaluating its potential use as both predictive biomarker and a therapeutic target.(36) Continuation of efforts using organoid models of PDAC to understand the mechanistic underpinnings of immunomodulating therapies and
  • 42. 36 advance research in pancreatic cancer early detection and precision medicine should accelerate improvement of patient outcomes for this deadly disease. VI. REFERENCES 1. Rahib L, Smith BD, Aizenberg R, Rosenzweig AB, Fleshman JM, and Matrisian LM. Projecting cancer incidence and deaths to 2030: the unexpected burden of thyroid, liver, and pancreas cancers in the United States. Cancer Res. 2014;74(11):2913-21. 2. Kleeff J, Korc M, Apte M, La Vecchia C, Johnson CD, Biankin AV, et al. Pancreatic cancer. Nat Rev Dis Primers. 2016;2:16022. 3. He J, Ahuja N, Makary MA, Cameron JL, Eckhauser FE, Choti MA, et al. 2564 resected periampullary adenocarcinomas at a single institution: trends over three decades. HPB (Oxford). 2014;16(1):83-90. 4. Oldfield LE, Connor AA, and Gallinger S. Molecular Events in the Natural History of Pancreatic Cancer. Trends Cancer. 2017;3(5):336-46. 5. Hruban RH, Goggins M, Parsons J, and Kern SE. Progression model for pancreatic cancer. Clin Cancer Res. 2000;6(8):2969-72. 6. Makohon-Moore A, and Iacobuzio-Donahue CA. Pancreatic cancer biology and genetics from an evolutionary perspective. Nat Rev Cancer. 2016;16(9):553-65. 7. Yachida S, Jones S, Bozic I, Antal T, Leary R, Fu B, et al. Distant metastasis occurs late during the genetic evolution of pancreatic cancer. Nature. 2010;467(7319):1114-7. 8. Notta F, Chan-Seng-Yue M, Lemire M, Li Y, Wilson GW, Connor AA, et al. A renewed model of pancreatic cancer evolution based on genomic rearrangement patterns. Nature. 2016;538(7625):378-82. 9. Tomasetti C, and Vogelstein B. Cancer etiology. Variation in cancer risk among tissues can be explained by the number of stem cell divisions. Science. 2015;347(6217):78-81. 10. Connor AA, Denroche RE, Jang GH, Timms L, Kalimuthu SN, Selander I, et al. Association of Distinct Mutational Signatures With Correlates of Increased Immune Activity in Pancreatic Ductal Adenocarcinoma. JAMA Oncol. 2017;3(6):774-83. 11. Waddell N, Pajic M, Patch AM, Chang DK, Kassahn KS, Bailey P, et al. Whole genomes redefine the mutational landscape of pancreatic cancer. Nature. 2015;518(7540):495-501. 12. Collisson EA, Sadanandam A, Olson P, Gibb WJ, Truitt M, Gu S, et al. Subtypes of pancreatic ductal adenocarcinoma and their differing responses to therapy. Nat Med. 2011;17(4):500-3. 13. Bailey P, Chang DK, Nones K, Johns AL, Patch AM, Gingras MC, et al. Genomic analyses identify molecular subtypes of pancreatic cancer. Nature. 2016;531(7592):47- 52. 14. Moffitt RA, Marayati R, Flate EL, Volmar KE, Loeza SG, Hoadley KA, et al. Virtual microdissection identifies distinct tumor- and stroma-specific subtypes of pancreatic ductal adenocarcinoma. Nat Genet. 2015;47(10):1168-78. 15. Noll EM, Eisen C, Stenzinger A, Espinet E, Muckenhuber A, Klein C, et al. CYP3A5 mediates basal and acquired therapy resistance in different subtypes of pancreatic ductal adenocarcinoma. Nat Med. 2016;22(3):278-87.
  • 43. 37 16. Farrell JJ, Elsaleh H, Garcia M, Lai R, Ammar A, Regine WF, et al. Human equilibrative nucleoside transporter 1 levels predict response to gemcitabine in patients with pancreatic cancer. Gastroenterology. 2009;136(1):187-95. 17. Jones S, Zhang X, Parsons DW, Lin JC, Leary RJ, Angenendt P, et al. Core signaling pathways in human pancreatic cancers revealed by global genomic analyses. Science. 2008;321(5897):1801-6. 18. Hanahan D, and Weinberg RA. Hallmarks of cancer: the next generation. Cell. 2011;144(5):646-74. 19. Biankin AV, Waddell N, Kassahn KS, Gingras MC, Muthuswamy LB, Johns AL, et al. Pancreatic cancer genomes reveal aberrations in axon guidance pathway genes. Nature. 2012;491(7424):399-405. 20. Kaukonen R, Mai A, Georgiadou M, Saari M, De Franceschi N, Betz T, et al. Normal stroma suppresses cancer cell proliferation via mechanosensitive regulation of JMJD1a- mediated transcription. Nat Commun. 2016;7:12237. 21. Rath N, and Olson MF. Regulation of pancreatic cancer aggressiveness by stromal stiffening. Nat Med. 2016;22(5):462-3. 22. Ansari D, Carvajo M, Bauden M, and Andersson R. Pancreatic cancer stroma: controversies and current insights. Scand J Gastroenterol. 2017;52(6-7):641-6. 23. Neesse A, Algul H, Tuveson DA, and Gress TM. Stromal biology and therapy in pancreatic cancer: a changing paradigm. Gut. 2015;64(9):1476-84. 24. Lunardi S, Jamieson NB, Lim SY, Griffiths KL, Carvalho-Gaspar M, Al-Assar O, et al. IP-10/CXCL10 induction in human pancreatic cancer stroma influences lymphocytes recruitment and correlates with poor survival. Oncotarget. 2014;5(22):11064-80. 25. Ene-Obong A, Clear AJ, Watt J, Wang J, Fatah R, Riches JC, et al. Activated pancreatic stellate cells sequester CD8+ T cells to reduce their infiltration of the juxtatumoral compartment of pancreatic ductal adenocarcinoma. Gastroenterology. 2013;145(5):1121- 32. 26. Sherman MH, Yu RT, Engle DD, Ding N, Atkins AR, Tiriac H, et al. Vitamin D receptor-mediated stromal reprogramming suppresses pancreatitis and enhances pancreatic cancer therapy. Cell. 2014;159(1):80-93. 27. Watt J, and Kocher HM. The desmoplastic stroma of pancreatic cancer is a barrier to immune cell infiltration. Oncoimmunology. 2013;2(12):e26788. 28. Poschke I, Faryna M, Bergmann F, Flossdorf M, Lauenstein C, Hermes J, et al. Identification of a tumor-reactive T-cell repertoire in the immune infiltrate of patients with resectable pancreatic ductal adenocarcinoma. Oncoimmunology. 2016;5(12):e1240859. 29. Balachandran VP, Luksza M, Zhao JN, Makarov V, Moral JA, Remark R, et al. Identification of unique neoantigen qualities in long-term survivors of pancreatic cancer. Nature. 2017;551(7681):512-6. 30. Vonderheide RH, and Bayne LJ. Inflammatory networks and immune surveillance of pancreatic carcinoma. Curr Opin Immunol. 2013;25(2):200-5. 31. Evans RA, Diamond MS, Rech AJ, Chao T, Richardson MW, Lin JH, et al. Lack of immunoediting in murine pancreatic cancer reversed with neoantigen. JCI Insight. 2016;1(14). 32. Hwang CI, Boj SF, Clevers H, and Tuveson DA. Preclinical models of pancreatic ductal adenocarcinoma. J Pathol. 2016;238(2):197-204.
  • 44. 38 33. Sato T, Vries RG, Snippert HJ, van de Wetering M, Barker N, Stange DE, et al. Single Lgr5 stem cells build crypt-villus structures in vitro without a mesenchymal niche. Nature. 2009;459(7244):262-5. 34. Boj SF, Hwang CI, Baker LA, Chio, II, Engle DD, Corbo V, et al. Organoid models of human and mouse ductal pancreatic cancer. Cell. 2015;160(1-2):324-38. 35. Tiriac H, Belleau P, Engle DD, Plenker D, Deschenes A, Somerville TDD, et al. Organoid Profiling Identifies Common Responders to Chemotherapy in Pancreatic Cancer. Cancer Discov. 2018;8(9):1112-29. 36. Guo X, Hollander L, MacPherson D, Wang L, Velazquez H, Chang J, et al. Inhibition of renalase expression and signaling has antitumor activity in pancreatic cancer. Sci Rep. 2016;6:22996. 37. Patra KC, Bardeesy N, and Mizukami Y. Diversity of Precursor Lesions For Pancreatic Cancer: The Genetics and Biology of Intraductal Papillary Mucinous Neoplasm. Clin Transl Gastroenterol. 2017;8(4):e86. VII. FIGURES Figure 1. Genetic evolution of pancreatic cancer. Pancreatic cancer may arise from either the development and progression of intraductal papillary mucinous neoplasm (top) or pancreatic intraepithelial neoplasm (bottom) as a result of sequential accumulation of characteristic driver mutations. This illustration was adapted from REF 37, with permission.
  • 45. 39 Figure 2. Creation of immunogenic murine organoid models of PDAC using KP-NINJA mouse model. (A) Schematic representation of major steps involved in the isolation of murine pancreatic organoids. (B) Genetic features of KP-NINJA mouse model for Cre-recombinase inducible mutation of KRAS and deletion of P53 (top), and multilayered control of inducible expression of GFP-tagged T cell neoantigens by Cre recombinase, rtTA-doxycycline and FLPo-
  • 46. 40 tamoxifen (bottom). GFP, green fluorescent protein; rtTA, reverse tetracycline-controlled transactivator; FLPo, codon-improved flippase recombinase. Figure 3. In vitro transformed murine pancreatic organoids recapitulates features of early PDAC in mouse. (A) Neoplastic transformation of murine pancreatic organoids by lentivirus encoding rtTA-Cre-iRFP670. Fluorescence imaging confirms RFP labeling of transformed cells in organoids. (B) Flow cytometry analysis confirms RFP expression in transformed organoids, which were subsequently sorted for the brightest 10% of cells expressing RFP. Leaky expression
  • 47. 41 of GFP-tagged neoantigens is not observed in these organoids. Untransformed organoids were used as a negative control. (C) Subcutaneous injection of transformed versus untransformed organoids in opposite flanks of mouse results in tumor formation only with transformed organoids. (D) H&E of tumor derived from subcutaneous injection transformed organoids in mouse. (E) H&E of primary pancreatic tumor from KP-C mouse model. (F) H&E of tumor derived from subcutaneous injection of PDAC cell lines generated from KP-C mouse model.
  • 48. 42 Figure 4. Modeling PDAC progression by serial in vivo transfer of transformed murine pancreatic organoids. (A) Experimental design for serial in vivo transfer of organoids. (B) H&E of organoid-derived tumors after successive rounds of in vivo transfer shows progressively more advanced tumors. (C) Immunohistochemical analysis of organoid-derived tumors after one round of in vivo transfer. (D) Fluorescence imaging of organoids reveals more uniform RFP labeling of organoids after one round of in vivo transfer versus organoids before in vivo transfer, indicating in vivo selection of transformed neoplastic organoids
  • 49. 43 Figure 5. High level of neoantigen expression in murine PDAC organoids results in rejection of tumor growth in mouse. (A) Experimental design for expression of GFP-tagged neoantigens in PDAC organoids by adenovirus encoding FLPo. (B) Flow cytometry analysis confirms GFP expression in organoids treated with adenovirus encoding FLPo, which were subsequently sorted for the brightest 10% of GFP-positive cells. Organoids treated with adenovirus encoding Cre was used as a negative control. (C) Subcutaneous injection of neoantigen-positive versus neoantigen-negative transformed organoids. First cohort of mice received neoantigen-negative organoids (N=5). Second cohort received neoantigen-positive organoids (N=6). Third cohort received neoantigen-positive organoids plus retroorbital injections of luciferase-positive P14 CD8+ T cells 24 hours prior to organoid injections (N=3). In vivo imaging after 24 hours of organoid injections reveals accumulation of luciferase-positive T cells at the site of organoid injections in the third cohort. (D) Mice were monitored for growth of tumors for up to 30 days. Tumor growth was observed in all of the mice in the first cohort. There was no tumor growth in any mouse in the second or third cohort that received neonantigen- positive organoids.
  • 50. 44 Figure 6. Low level of neoantigen expression in murine PDAC organoids permits tumor growth with increased immune infiltration. (A) A murine PDAC organoid line that expresses GFP-tagged T cell neoantigens at a low level was generated by dilution of neoantigen-positive organoids with neoantigen-negative organoids. Flow cytometry analysis confirms GFP expression in only 10% of the total population. (B) Subcutaneous injection of organoids generated from (A) resulted in growth of tumors in mouse. H&E of tumors derived from these organoids shows increased immune infiltration compared to tumors derived from neoantigen- negative organoids.
  • 51. 45 Figure 7. Development of co-culture model system for murine PDAC organoids and T cells. (A) P14 CD8+ T cells were sorted from splenocytes of P14 mouse following in vitro expansion and pre-activation with IL-2 and GP33 peptide, respectively. Blue calcein dye was used to label
  • 52. 46 prepared T cells and green calcein dye was used to label both T cells and PDAC organoids. T cells were introduced to the liquid medium of wells containing either neoantigen-positive or neoantigen-negative PDAC organoids which were maintained in Matrigel plugs in 24-well plate format. Fluorescence imaging of co-cultures at 1 hour demonstrates prominent clustering of T cells at the boundaries of Matrigel plugs. (B) Fluorescence imaging of co-cultures at 24 hours reveals evidence of increased T cell infiltration of Matrigel plugs containing neoantigen-positive organoids. Images were converted to black and white for better visualization of blue dye. IL, interleukin; GP, glycoprotein.
  • 53. 47 Figure 8. Human PDAC organoids form tumors in mouse that are histologically matched to patient-derived primary tissues. (A) Schematic overview for the creation of patient-derived organoids using different types of primary tissues, including EUS-FNB specimens, surgical resection specimens, and tissues from PDX mouse models that were established by implanting a piece of surgical resection specimen in mouse. (B) Patient-derived organoids validated by Sanger sequencing of organoids for mutations in KRAS and P53, in vivo transfer of organoids for tumor
  • 54. 48 formation in mouse, and IHC analysis of organoid-derived tumors. (C) IHC analyses of tumors generated from different types of primary tissues are shown. Row X shows a tumor generated from organoids derived from FNB. Row Y shows a tumor directly taken from a PDX mouse model. Row Y’ shows a tumor generated from organoids derived from the tumor shown in row Y. Tumors shown in rows Y and Y’ were ultimately derived from the same patient and are histologically matched. (D) H&E of three additional tumors derived from FNB specimens are shown. Bx120817 (left) was derived from the primary tumor of a patient who had borderline resectable disease. Bx111417 (middle) was derived from the primary tumor of a patient who had metastatic disease. Bx102417 (right) was derived from a metastatic lesion in the liver. (E) Sanger sequencing of patient-derived organoids reveals classic G12V mutation in KRAS. PDX060917 (left) was derived from tissues from a PDX mouse model and Bx011218A (right) was derived from an EUS-FNB specimen. EUS-FNB, endoscopic ultrasound-guided fine-needle biopsy; PDX, patient-derived xenograft; IHC, immunohistochemistry. VIII. TABLES Mutated gene Frequency (%) Effect of mutation Cellular process or pathway affected Biological significance of mutation KRAS 95 Gain of function RAS–ERK pathway Ligand-independent cell proliferation and survival; immunosuppression; metabolic alterations CDKN2A 90 Loss of function G1/S transition G1/S checkpoint failure TP53 80-85 Gain of function DNA damage response G1/S checkpoint failure; G2/M checkpoint failure; apoptosis resistance SMAD4 55 Loss of function TGFβ pathway Failure of celluar homeostasis; loss of TGFβ- and TP53-mediated gene expression TGFBR1 ≤10 Loss of function TGFβ pathway Failure of celluar homeostasis; loss of TGFβ- and TP53-mediated gene expression
  • 55. 49 TGFBR2 ≤10 Loss of function TGFβ pathway Failure of celluar homeostasis; loss of TGFβ- and TP53-mediated gene expression ARID1A ≤10 Loss of function Epigenomic reprogramming - SWI/SNF Loss of regulatory function in modulating nucleosomal DNA- histone interactions ARID1B ≤10 Loss of function Epigenomic reprogramming - SWI/SNF Loss of regulatory function in modulating nucleosomal DNA- histone interactions ARID2 ≤10 Loss of function Epigenomic reprogramming - SWI/SNF Loss of regulatory function in modulating nucleosomal DNA- histone interactions KMT2C ≤10 Loss of function Epigenomic reprogramming - KMT2 Decreased methylation of H3K4 KMT2D ≤10 Loss of function Epigenomic reprogramming - KMT2 Decreased methylation of H3K4 KMT2A ≤10 Loss of function Epigenomic reprogramming - KMT2 Decreased methylation of H3K4 SF3B1 ≤10 Altered function RNA splicing Loss of polycomb repressive complex-mediated transcriptional regulation of HOX genes; abnormal splicing of pre-mRNA PCDH15 ≤10 Loss of function Homophilic cell adhesion Disruption of cadherin-mediated calcium-dependent cell adhesion BRAF ≤5 Gain of function RAS–ERK pathway Ligand-independent cell proliferation and survival; immunosuppression; metabolic alterations APC2 ≤5 Loss of function G1/S transition G1/S checkpoint failure CHD1 ≤5 Loss of function G1/S transition G1/S checkpoint failure FBXW7 ≤5 Loss of function G1/S transition G1/S checkpoint failure ATM ≤5 Loss of function DNA damage response G1/S checkpoint failure; G2/M checkpoint failure; apoptosis resistance ACVR1B ≤5 Loss of function TGFβ pathway Failure of celluar homeostasis; loss of TGFβ- and TP53-mediated gene expression SMAD3 ≤5 Loss of function TGFβ pathway Failure of celluar homeostasis; loss of TGFβ- and TP53-mediated gene expression
  • 56. 50 PBRM1 ≤5 Loss of function Epigenomic reprogramming - SWI/SNF Loss of regulatory function in modulating nucleosomal DNA- histone interactions SMARCA2 ≤5 Loss of function Epigenomic reprogramming - SWI/SNF Loss of regulatory function in modulating nucleosomal DNA- histone interactions SMARCA4 ≤5 Loss of function Epigenomic reprogramming - SWI/SNF Loss of regulatory function in modulating nucleosomal DNA- histone interactions MKK4 ≤5 Loss of function Cellular stress response Failure of JNK signaling; disruption of TLR signaling ROBO1 ≤5 Loss of function Axon guidance Abnormal migration of cells ROBO2 ≤5 Loss of function Axon guidance Abnormal migration of cells SLIT ≤5 Loss of function Axon guidance Abnormal migration of cells Table 1. Mutational landscape of pancreatic cancer. Commonly mutated genes in PDAC are organized by frequency of mutation in PDAC, effect of mutation on gene function, celluar process or signaling pathway affected, and biological significance of mutation. This table is a summary of data described in greater detail in REF 6.