0% found this document useful (0 votes)
345 views291 pages

1138487627

Uploaded by

Eugen M
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
345 views291 pages

1138487627

Uploaded by

Eugen M
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 291

JAK-STAT Signaling in

Diseases

JAK-STAT Signaling in

Diseases

Edited by

Ritobrata Goswami

CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway, Suite 300
Boca Raton, FL 33487–2742
© 2020 by Taylor & Francis Group, LLC
CRC Press is an imprint of Taylor & Francis Group, an Informa business
No claim to original U.S. Government works
International Standard Book Number-13: 978-1-138-48762-8 (Hardback)
This book contains information obtained from authentic and highly regarded sources. Reasonable efforts have been
made to publish reliable data and information, but the author and publisher cannot assume responsibility for the
validity of all materials or the consequences of their use. The authors and publishers have attempted to trace the
copyright holders of all material reproduced in this publication and apologize to copyright holders if permission to
publish in this form has not been obtained. If any copyright material has not been acknowledged, please write and
let us know so we may rectify in any future reprint.
Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or
utilized in any form by any electronic, mechanical, or other means, now known or hereafter invented, including
photocopying, microfilming, and recording, or in any information storage or retrieval system, without written
permission from the publishers.
For permission to photocopy or use material electronically from this work, please access www.copyright.com (www.
copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923,
978-750-8400. CCC is a not-for-profit organization that provides licenses and registration for a variety of users. For
organizations that have been granted a photocopy license by the CCC, a separate system of payment has been
arranged.
Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used only for
identification and explanation without intent to infringe.

Library of Congress Control Number: 2020932394

Visit the Taylor & Francis Web site at


www.taylorandfrancis.com
and the CRC Press Web site at
www.crcpress.com
Contents

Preface........................................................................................................................ vii

Editor.......................................................................................................................... ix

Contributors ................................................................................................................. xi

1. Regulation of Cytokine Signaling by the JAK-STAT Pathway ........................................... 1

Nicolette Nadene Houreld

2. The Structure-Function Bonhomie of JAK-STAT Molecules.............................................. 9

Ritobrata Goswami

3. MicroRNA-Mediated Regulation of JAK-STAT Signaling in Non-Cancerous

Human Diseases ..................................................................................................... 35

Chandra S. Boosani, Wanlin Jiang, Taylor Burke, and Devendra K. Agrawal

4. JAK-STAT Signaling in Asthma and Allergic Airway Inflammation.................................. 49

Amina Abdul Qayum, Tristan Hayes and Mark H. Kaplan

5. Role of JAK-STAT Signaling in Atopic Dermatitis........................................................ 69

Radomir M. Slominski and Matthew J. Turner

6. JAK-STAT Signaling Pathway and Gliosis in Neuroinflammatory Diseases........................ 83

Han-Chung Lee, Kai-Leng Tan, Pike-See Cheah and King-Hwa Ling

7. JAK-STAT Signaling in Cardiovascular Disease ..........................................................103

George W. Booz, Raffaele Altara and Sean P. Didion

8. Diabetes and Obesity: Abnormal JAK-STAT Signaling .................................................123

Marcia J. Abbott

9. JAK-STAT Signaling in Liver Fibrosis.......................................................................143

Marwa K. Ibrahim and Noha G. Bader El Din

10. Renal Disorders: Involvement of JAK-STAT Pathway...................................................159

Yuji Nozaki

11. JAK-STAT Signaling in Hematologic Malignancies .....................................................177

Thomas Pincez and Thai Hoa Tran

12. Aberrant JAK-STAT Signaling in Hematopoietic Malignancies ......................................225

Parvis Sadjadian

v
vi Contents

13. Immunodeficiency: Consequences of Mutations in JAK-STAT Signaling ...........................241

Daniel Silberger and Duy Pham

14. Targeting JAK-STAT Pathway for Various Inflammatory Diseases and

Viral Infections ......................................................................................................257

Christina Gavegnano and Raymond F. Schinazi

Index .................................................................................................................. 269

Preface

JAK-STAT (Janus kinase-Signal Transducers and Activators of Transcription) signaling is one of


the few conserved signaling cascades transmitting multiple signals required for homeostasis in
animals and humans. In JAK-STAT signaling, extracellular ligand binding to its cognate receptor
leads to pathway activation and signal transduction. This pathway is critical for hematopoiesis,
immune system development, and other events. Activation of JAK leads to induction of cell
proliferation, cell migration, differentiation, and apoptosis. Mutations in JAK-STAT pathway can
either lead to inflammatory diseases or may impede regular homeostasis.
JAK-STAT pathways can cross-talk with other signaling pathways. JAK-STAT pathway has
been targeted to develop drugs to downregulate the immune response. Few of the JAK inhibitors
are in Phase III of clinical trial and have been approved by FDA. Tofacitinib that targets JAK3
has shown its effect to treat rheumatoid arthritis, psoriasis. OPB-51602, which targets STAT3, is
in phase I to treat nasopharyngeal carcinoma. JAK inhibitors have an advantage over cytokine
receptor blocking drugs in that they that can be taken orally as they are small molecule drugs.
JAK-STAT pathway is a very well-studied signaling pathway. The book has consolidated
information on the role of JAK-STAT signaling in autoimmune and inflammatory diseases
including asthma, atopic dermatitis, hepatic fibrosis, diabetes and obesity, cancer, cardiovascular
disease, immunodeficiency, renal disorders, neuro-inflammatory disorders, among others. This
book is unique as it encompasses different aspects of JAK-STAT signaling and how dysregulation
of this signaling pathway is associated with various disorders that are prevalent worldwide leading
to morbidity and mortality associated with increased expenditure on healthcare and loss of
manpower. This book has an update on the therapeutic implication of JAK-STAT inhibitors
different phases of clinical trial. I sincerely believe this book will stimulate interest in graduate
students, academicians and scientific community. This book will be good reference point for
industry professionals who are involved in translational research leading to develop products to
various diseases. Positive feedback of this book from the readers will allow us to expand the
disease repertoire that is associated with dysregulated JAK-STAT signaling.
Editor

Ritobrata Goswami, PhD, is presently working as an Assistant Professor in the


School of Bioscience, IIT Kharagpur since 2016. Prior to his present affiliation,
Dr Goswami was associated with the Division of Biological and Life Sciences,
Ahmedabad University (2014-16). He received his bachelor (2005) and mas­
ters’ degree (2007) in Biotechnology from West Bengal University of Tech­
nology and Birla Institute of Technology & Science, Pilani; respectively.
Dr Goswami obtained his PhD at Indiana University, Indianapolis, USA
(2012) followed by post-doctoral training at Icahn School of Medicine at
Mount Sinai, New York, USA (2012-14). He is a life member of the Indian
Immunology Society. Dr Goswami has extramural projects funded by the Government of India.
Dr Goswami’s research interests include the role of nutrients and hormones in the development
and function of T helper cells to regulate inflammatory disorders. He is keen on identifying
a regulating network of transcription factors involved in autoimmune and inflammatory disorders so
that better therapeutics can be developed.

ix
Contributors

Marcia J. Abbott Pike-See Cheah


Department of Health Sciences Department of Human Anatomy, Faculty of
Crean College of Health and Behavioral Medicine and Health Sciences
Sciences, Chapman University Universiti Putra Malaysia
Orange, California Seri Kembangan, Malaysia

Devendra K. Agrawal Sean P. Didion


Department of Clinical and Translational Science Department of Pharmacology and Toxicology
Creighton University School of Medicine The University of Mississippi Medical Center
Omaha, Nebraska Jackson, MS

Raffaele Altara Christina Gavegnano


Institute for Experimental Medical Research Center for AIDS Research, Laboratory of
Oslo University Hospital and University of Biochemical Pharmacology, Department of
Oslo Pediatrics
Oslo, Norway Emory University
Atlanta, GA
Noha G. Bader El Din
Department of Microbial Biotechnology Ritobrata Goswami
Genetic Engineering and Biotechnology School of Bioscience
Division Indian Institute of Technology Kharagpur
National Research Centre Kharagpur, India
Dokki, Egypt
Tristan Hayes
Chandra S. Boosani Department of Pediatrics and Herman B Wells
Department of Clinical and Translational Center for Pediatric Research
Science Indiana University School of Medicine
Creighton University School of Medicine Indianapolis, Indiana
Omaha, Nebraska
Nicolette Nadene Houreld
George W. Booz Laser Research Centre, Faculty of
Department of Pharmacology and Toxicology Health Sciences
The University of Mississippi University of Johannesburg
Medical Center Johannesburg, South Africa
Jackson, MS
Marwa K. Ibrahim
Taylor Burke Department of Microbial Biotechnology,
Department of Clinical and Translational Genetic Engineering and Biotechnology
Science Division
Creighton University School of Medicine National Research Centre
Omaha, Nebraska Dokki, Egypt

xi
xii Contributors

Wanlin Jiang Parvis Sadjadian


Department of Clinical and Translational University Clinic for Hematology, Oncology,
Science Hemostaseology and Palliative Care,
Creighton University School of Medicine Johannes Wesling Medical Center
Omaha, Nebraska Minden
University of Bochum
Mark H. Kaplan Minden, Germany
Department of Pediatrics and Herman B Wells
Center for Pediatric Research Raymond F. Schinazi
Indiana University School of Medicine Center for AIDS Research, Laboratory of
Indianapolis, Indiana Biochemical Pharmacology
Department of Pediatrics, Emory
Han-Chung Lee University
Laboratory Centre Atlanta, GA
Xiamen University Malaysia
Sepang, Malaysia Daniel Silberger
Department of Pathology
King-Hwa Ling University of Alabama at
Department of Biomedical Sciences, Birmingham
Faculty of Medicine and Health Sciences Birmingham, Alabama
Universiti Putra Malaysia
Seri Kembangan, Malaysia Radomir M. Slominski
Department of Dermatology
Yuji Nozaki Indiana University School of
Department of Hematology and Medicine
Rheumatology Indianapolis, Indiana
Kindai University School of Medicine
Osaka-Sayama, Japan Kai-Leng Tan
Institute of Biomedical and Pharmaceutical
Duy Pham Sciences
Department of Pathology Guangdong University of Technology
University of Alabama at Birmingham Guangzhou, China
Birmingham, Alabama
Thai Hoa Tran
Thomas Pincez Division of Pediatric
Division of Pediatric Hematology-Oncology Hematology-Oncology
Charles-Bruneau Cancer Center, Ste-Justine Charles-Bruneau Cancer Center,
Hospital Ste-Justine Hospital
Quebec, Canada Quebec, Canada

Amina Abdul Qayum Matthew J. Turner


Department of Pediatrics and Herman B Wells Department of Dermatology
Center for Pediatric Research Indiana University School of
Indiana University School of Medicine Medicine
Indianapolis, Indiana Indianapolis, Indiana
1
Regulation of Cytokine Signaling by the
JAK-STAT Pathway

Nicolette Nadene Houreld


Laser Research Centre, Faculty of Health Sciences
University of Johannesburg
Johannesburg, South Africa

1.1 Introduction
The Janus kinase/signal transducers and activators of transcription (JAK-STAT) pathway is a prompt
pleiotropic cytoplasmic to nuclear signaling pathway used to transduce a variety of signals, activated
by cytokines, hormones, and growth factors, for development and homeostasis. The JAK-STAT
pathway is responsible for controlling signals of over fifty cytokines, growth factors, and hormones
(Morris, Kershaw, and Babon 2018; Hammaren et al. 2019), while negative regulation is through
suppressor of cytokine signaling (SOCS) proteins (bind to and inactivate JAK3), and the protein
inhibitors of activated STATs (PIAS; bind to STAT dimers thereby preventing DNA binding).
Cytokines are glycoproteins (ligands) secreted by cells and operate as intercellular messengers,
inducing differentiation, proliferation, growth, and apoptosis of their target cells.
Signaling via the JAK-STAT pathway is instigated by binding of a ligand to its receptor. Binding
results in dimerization, oligomerization, and/or conformational changes of the receptor complex,
which allow JAK proteins to bind to the receptor complex intracellular domains inducing trans­
autophosphorylation of the tyrosine residues (JH1), converting the receptor into a tyrosine kinase.
Phosphorylated chains serve as docking sites for SH2 domain-containing signaling molecules such as
STATs. Receptor-bound STATs are phosphorylated by JAK on a specific tyrosine in the C-terminal
tail, allowing them to form homo- and heterodimers, which rapidly translocate into the nucleus. In the
nucleus, they associate with proteins and produce transcriptional complexes/factors with extensive
effects on regulation of transcription and epigenetics (Hammaren et al. 2019).

1.2 The JAKs


Janus (kinase) comes from the Roman mythological two-faced god, who looks to the future and
the past. JAK relates to the two faces due to the presence of two kinase domains, namely the
pseudokinase domain (JAK homology 2, JH2) and a catalytically active kinase domain (JH1).
They also contain an N-terminal band, and a four-point-one, ezrin, radixin, moesin (FERM)­
domain, which mediate the interaction of JAKs with their receptors. JAKs combine with the
proline-rich, membrane-proximal box1/box2 domain on cytokine receptors. They also contain an
Src homology 2 (SH2)-like domain, of unknown function, which lies between the pseudokinase
and FERM domains (Figure 1.1a) (Schindler, Levy, and Decker 2007; Hammaren et al. 2019).
JH2 has significant regulatory functions and is a source of numerous mutations that is the cause
of various diseases and disorders, including hematopoietic malignancies (JAK2 mutations),

1
2 JAK-STAT Signaling in Diseases

FIGURE 1.1 Structural organisation of JAK-STAT proteins. (a) JAKs share seven conserved homology domains:
JH1 serves as the catalytic, kinase domain, while JH2 represents the pseudokinase domain. JH3 and half of JH4
include the nonfunctional SH2-domain, and half of JH4 to JH7 includes a FERM domain. (b) STATs share seven
domains: the amino-terminal domain (NH2), the coiled-coil domain, the DNA-binding domain, the linker domain, the
SH2 domain, the tyrosine-activation domain, and the transactivation domain.

leukemia and lymphomas (all JAKs), and cancer (JAK1, JAK3) (Hammaren et al. 2019). There
are four JAKs in mammals (JAK1, JAK2, JAK3, and TYK2). JAK1, JAK2, and TYK2 are
ubiquitously expressed and relatively constitutive in their expression, while the expression of
JAK3 is mostly confined to cells of hematopoietic origin, and its expression is more inducible.
JAK1 associates with type I (IFN-α/β), type II (IFN-γ), IL-2, and IL-6 receptors. JAK2
interacts with single-chain receptors (i.e., EPOR, GH-R, and PRL-R) and IL-3 (IL-3R, IL-5R,
and GM-CSFR) cytokine families, as well as the IFN-γ receptor. Leukocyte-specific JAK3
exclusively associates with the IL-2 receptor γ-chain, and Tyk2 associates with receptors for
IFN-I, IL-6, IL-10, and IL-12/23 cytokine families (Schindler, Levy, and Decker 2007).

1.3 The STATs


The STAT family includes STAT1, STAT2, STAT3, STAT5A/B, and STAT6. STAT proteins
consist of seven well-defined, conserved domains: an N-terminal conserved domain (NH2, critical
for STAT function); a coiled-coil domain (involved in receptor binding, and associates with
regulatory proteins); a DNA-binding domain (DBD, cooperate in binding to the promoters of
target genes); a linker region (LK, spacer to maintain proper conformation between the dimeriza­
tion and DNA binding domains); an SH2 domain (critical for the recruitment of STATs to
activated receptor complexes and for the interaction with JAK and Src kinases); a tyrosine-
activation domain (Y); and a C-terminal transactivation domain (TAD, modulates the transcrip­
tional activation of target genes and vary considerably among STAT family members) (Jatiani
et al. 2010) (Figure 1.1b).

1.4 Cytokine Receptors


Cytokines function by binding to their associated transmembrane receptor, which triggers
intracellular signaling events and pathways that result in the alteration of gene expression. Most
Cytokine Signaling Regulation by JAK-STAT 3

of these receptors consist of a unique ligand-binding subunit and a signal-transducing subunit.


Often the signal transducing or cytoplasmic subunits are structurally similar to other cytokine
receptors, particularly in regions labeled as box 1 or the proline-rich motif and the box-2 motif,
and this is critical for proper receptor functioning and mediating of mitogenic signals. The ligand-
binding subunit, or membrane distal region, remains uniquely different to ensure differentiation
(Jatiani et al. 2010). Cytokine binding results in receptor tyrosine phosphorylation.
Cytokine receptors are divided into type I and type II receptors. Type I cytokine receptors bind
to and react to cytokines with four α-helical strands and share an amino acid motif (WSXWS).
Type II cytokine receptors are similar to type I, but lack the WSXWS motif. Cytokine receptors
signal through the JAK-STAT pathway and other pathways that typically trigger activation of the
mitogen-activated protein (MAP) kinase cascade. Different types of cells and tissues express well-
defined and diverse receptor combinations that respond to cytokine combinations unique to their
microenvironment. Thus, at any particular time, a single cell may respond to signals from multiple
cytokine receptors (Murray 2007). Different receptor classes preferentially associate with one JAK
family member, or a JAK combination.
Typically, receptors required for hematopoietic cell development and proliferation prefer JAK2;
common γ-chain receptors utilize JAK1 and JAK3, while other receptors use only JAK1 (Murray
2007). All interferons (IFNs), which are essential mediators of innate immunity against bacterial
and viral infection, as well as the interleukin(IL)-10 family (IL-10, IL-19, IL-20, IL-22, IL-24, IL­
26), anti-inflammatory cytokines, function through type II cytokine receptors, which dimerize in
multiple combinations to generate distinct downstream effects. JAK1 is imperative for signaling
through these type II receptor complexes (Ferrao et al. 2016).

1.5 Activation of JAK-STAT Pathways by Cytokines


Cytokine signaling through the JAK-STAT pathway regulates numerous cellular responses, including
proliferation, differentiation, motility, and cell survival. JAKs mediate signaling of around fifty to
sixty different hormones, cytokines, and growth factors ranging from regulators of the immune
system and hematopoiesis, such as IFNs, ILs, thrombopoietin (TPO), and erythropoietin (EPO), to
regulators of development and metabolism, such as growth hormone (GH) and prolactin (PRL)
(Table 1.1) (O’Shea and Plenge 2012).
Ligand binding results in receptor dimerization/oligomerization, which in turn results in the
juxtapositioning of JAKs through homodimeric or heterodimeric interactions. This leads to
the autophosphorylation and/or transphosphorylation of JAK, which causes the phosphoryla­
tion of receptor target tyrosine residues that serve as docking sites and allow the binding of
other SH2 domain–containing signaling molecules such as STAT (Jatiani et al. 2010)
(Figure 1.2). In unstimulated cells, STATs are idle, unphosphorylated cytosolic proteins.
Cellular stimulation via cytokines induces phosphorylation of receptor tyrosine residues that
serve as docking sites for STATs via their SH2 domains. Once phosphorylated, dimerized
STATs translocate to the nucleus and drive the transcription of cytokine inducible genes.
STAT activation is swift, with a maximum accumulation of phosphorylated STAT1 in the
nucleus within 30 min (Lim and Cao 2006).
Recently, a number of biophysical and protein engineering studies have presented new data,
which emphasizes the intricacy and complexity of signaling sparked by a cytokine–cytokine
receptor complex. This permits cytokines to produce diverse biological responses in spite of
using a marginal set of surface receptors and effector signaling molecules (Gorby et al. 2018).
Cytokine receptors will interact with the JAK-STAT pathway though its interaction with a
combination of JAKs. Cytokine-receptor interactions result in various JAK activation that exists
in association with cytokine receptors, for which activation is essential for the activation of
STATs. On the other hand, activated JAKs do not seem to exhibit specificity for a particular
STAT, as different receptors stimulate a common STAT even though they activate distinctively
4 JAK-STAT Signaling in Diseases

TABLE 1.1
Type I and Type II Cytokine Receptors and their Corresponding JAKs and STATs. Associated JAK­
STAT Proteins for which the Data is Weaker are Shown in Brackets (adapted from (Hammaren et al.
2019)).
Cytokine JAKs STATs

Type I Cytokine IL-6 JAK1, JAK2, TYK2 STAT3, STAT1


Receptors IL-11 JAK1, JAK2, TYK2 STAT3, STAT1
LIF JAK1, JAK2, TYK2 STAT3, STAT1
CNTF JAK1, (JAK2, TYK2) STAT3, (STAT1)
CLCF1 JAK1, (JAK2) STAT3, STAT1
CT-1 JAK1, (JAK2, TYK2) STAT3
OSM JAK1, (JAK2, TYK2) STAT3, STAT1
IL-31 JAK1, (JAK2) STAT3, STAT5,STAT1
G-CSF JAK1, (JAK2) STAT3
Leptin JAK2 STAT3
IL-12 TYK2, JAK2 STAT4
IL-23 TYK2, JAK2 STAT3, STAT4, STAT1
IL-27 JAK1, JAK2 STAT1, STAT3, STAT4, (STAT5)
IL-35 JAK1, JAK2 STAT1, STAT4
IL-2 JAK1, JAK3, (JAK2) STAT5, (STAT3)
IL-4 JAK1, JAK3 STAT6
IL-7 JAK1, JAK3 STAT5, (STAT3)
IL-9 JAK1, JAK3 STAT5, STAT3
IL-15 JAK1, JAK3 STAT5, (STAT3)
IL-21 JAK1, JAK3 STAT3, STAT5, (STAT1)
TSLP JAK1, JAK2 STAT1, STAT3, STAT4, STAT5, STAT6
IL-13 JAK1, JAK2, TYK2 STAT6, (STAT3)
IL-3 JAK2, JAK1 STAT5, STAT3
IL-5 JAK2 STAT5, STAT1, STAT3
GM-CSF JAK2 STAT5
EPO JAK2 STAT5
GH JAK2 STAT5, (STAT3)
PRL JAK2 STAT5
TPO JAK2 STAT5
Type II Cytokine IFN-I JAK1, TYK2 STAT1, STAT2, STAT3, STAT4,
Receptors (type I) (STAT5, STAT6)
IFN-γ JAK1, JAK2 STAT1
(type II)
IL-28a, JAK1, TYK2 STAT1, STAT2, STAT3, STAT5
IL-28b
IL-29 JAK1, TYK2 STAT1, STAT2, STAT3, STAT5
IL-10 JAK1, TYK2 STAT3, STAT1
IL-19 JAK1, JAK2 STAT3, STAT1
IL-20 JAK1, JAK2 STAT3, STAT1
IL-24 JAK1, JAK2 STAT3, STAT1
IL-22 JAK1, TYK2 STAT3, STAT1, (STAT5)
IL-26 JAK1, TYK2 STAT3, STAT1
Interleukin (IL); Leukemia inhibitory factor (LIF); Ciliary neurotrophic factor (CNTF); Cardiotrophin-like cytokine
factor 1 (CLCF1); Cardiotrophin-1 (CT-1); Oncostatin M (OSM); Granulocyte-colony stimulating factor (G-CSF);
Thymic stromal lymphopoietin (TSLP); Granulocyte-macrophage colony-stimulating factor (GM-CSF); Erythropoietin
(EPO); Growth hormone (GH); Prolactin (PRL); Thrombopoietin (TPO).
Cytokine Signaling Regulation by JAK-STAT 5

FIGURE 1.2 Overview of the JAK-STAT signaling pathway. Ligand binding (1) results in receptor dimerization, JAK
activation, and receptor phosphorylation (2). STAT binds to the phosphorylated receptor (3), which becomes
phosphorylated by JAK (4). STAT dimers form (5), which translocates into the nucleus and stimulates gene
transcription (6). Negative regulation of JAK-STAT signaling is through suppresser of cytokine signaling (SOCS)
proteins, which directly bind to and inactivate JAKs, and protein inhibitors of activated STATs (PIAS) that bind to
phosphorylated STAT dimers, averting DNA binding.

different JAKs. It would appear that the specificity for the activation of STAT lies with the STAT
docking sites on the receptors themselves (Jatiani et al. 2010).
Once activated STAT has translocated to the nucleus, it binds to a consensus DNA-recognition
motif known as gamma-activated sites (GAS) in the promoter region of cytokine-inducible genes and
activates transcription. These GAS-like elements act as ligands for a variety of STAT family members.
There are three key features of STAT-binding elements (SBE), which include the core motif (with
sequence TT-(X)n-AA; where X is a G/C rich palindromic sequence in length n), the core spacing
between the palindromic A/T residues, and composition of the sequence between palindromic A/T
residues. Generally, there is a preference for a specific core spacing for each class of cytokine­
responsive genes, for example, GAS elements consistently occupy five-base pair core spacing, whereas
genes that respond to IL-6 occupy a four-base core spacing (Wilks and Harpur 1994). STAT1 and
STAT5 prefer sites with a three-base pair spacer; however, STAT5 has also been shown to bind weakly
to a four base pair spacer. STAT6 prefers sites with a four-base pair spacer, and STAT4 prefers the
palindromic sequence (T/A)TTCC(C/G)GGAA(T/A) where the first and last T/A sites outside of the
usual motif are also necessary for binding (Morris, Kershaw, and Babon 2018).

1.6 JAK-STAT Regulation


A characteristic feature of JAK-STAT signaling is its rapid onset and subsequent decay. Activated
STATs accumulate within a few hours in the nucleus, however, the signal promptly decays and the
6 JAK-STAT Signaling in Diseases

STATs are re-exported back to the cytoplasm for the next round of signaling. This decline in
activity involves down-regulation of both receptors and JAK proteins, as well as STAT-transcriptional
activity (Schindler, Levy, and Decker 2007). Three well-defined mechanisms of signal decay
include dephosphorylation by protein phosphotyrosine phosphatases, nuclear export, and feed­
back inhibition via SOCS, as well as inhibition by PIAS. Post-translational modifications (PTMs)
of STAT proteins institute another important regulatory mechanism. STATS also undergo post-
translational covalent modifications such as ubiquitination (post-translational modification by
adding ubiquitin to the protein sequence).
Protein phosphotyrosine phosphatases (PTPs) negatively regulate the JAK-STAT pathway by
dephosphorylating tyrosine residues. There are six PTPs that regulate JAK-STAT. Receptor and
JAK dephosphorylation has been carried out by phosphatases, and include SHP-1, SHP-2, and
CD45, while STAT dephosphorylation is carried out by SHP-2, PTP1B, TC-PTP, and PTP-BL
(Schindler, Levy, and Decker 2007; Seif et al. 2017). Since PTPs are constitutively expressed, they
tend to restrain the amplitude of the signaling cascade rather than controlling its duration
(Morris, Kershaw, and Babon 2018). The process of nuclear translocation is complex and there
is a balance between STAT nuclear import and export. This seems to be controlled by multiple
nuclear export sequence (NES) and nuclear localization sequence (NLS) elements. During activa­
tion of cellular signaling, there is a shift toward STAT accumulation in the nucleus. This balance
is shifted toward nuclear export when there is signal decay (Schindler, Levy, and Decker 2007).
Proteolytic processing of STAT via cleavage of the C-terminal acts as a general mechanism for the
negative regulation of STAT protein function (Hendry and John 2004). This may occur in the
nucleus or in the cytoplasm, and this protease may cleave activated (phosphorylated) or inactivated
STAT. STAT are rapidly inactivated by dephosphorylation with a half-life of phosphorylated
STAT1 at less than 15 min (Lim and Cao 2006).
Post-translational modifications regulate every aspect of transcription factor function and
coordinate access of RNA polymerases to promoter templates. Site-specific, DNA-binding
transcription factors repress, activate, enhance, or silence complexes and associated enzymatic
activities (Filtz, Vogel, and Leid 2014). Cytokine signal transduction is dependent on the ability of
individual receptors to recruit signal transcription factors (STFs), which in turn activate the
transcription of the genes which are related to that particular cytokine (Wilks and Harpur 1994).
This process is very complex and relies on affinity differences of STFs for GAS elements. There
are two main subclasses of DNA response elements in gene promoters, and include interferon-
stimulated response element (ISRE) and GAS-like sequences. ISRE is located upstream of
interferon alpha/beta (IFN-α/β) inducible genes, while GAS is located upstream of IFN gamma
(IFN-γ)-inducible genes. GAS-like elements, which bind to STFs induced by other cytokines, all
consist of the same basic DNA motif with slight changes in individual genes (Wilks and Harpur
1994). Owing to activation of other transcription factors and crosstalk between different signal
transducing pathways, determination of the complete set of genes upregulated by STAT proteins
is difficult. However, it is known that several hundred to thousands of genes can be upregulated or
down-regulated in response to each STAT protein (Morris, Kershaw, and Babon 2018).
The SOCS family members are the primary drivers of signal attenuation and are negative-
feedback inhibitors of signaling induced by cytokines and other stimuli that act via the JAK­
STAT pathway. There are eight SOCS proteins: SOCS1 to 7 and CIS. These proteins all contain
an SH2 domain and a short, C-terminal domain, known as the SOCS box. The SOCS box is
associated with an adapter complex, elongin BC, which allows for the recruitment of an E3
ubiquitin ligase scaffold (Cullin5) to catalyse and ubiquitinate signaling intermediates recruited by
their SH2 domains (Liau et al. 2018). Ubiquitination is mediated by an ubiquitin-activating
enzyme (E1), ubiquitin-conjugating enzyme (E2), and ubiquitin ligase (E3), which mediates
between E2 and the substrate. The resulting poly-ubiquitinated conjugates are rapidly identified
and degraded by 26S proteasome. SOCS proteins form E3 ubiquitin ligase complexes with Cullin­
2, Elongin B, Elongin C, and Rbx1 and ubiquitinate JAK and receptor molecules, followed by
their degradation and internalization, respectively (Hatakeyama 2012). SOCS1 and SOCS3 are
capable of directly binding to and inhibiting the kinase activity of JAK, and is due to a short
Cytokine Signaling Regulation by JAK-STAT 7

motif, kinase inhibitory region (KIR), upstream of the SH2 domain. SOCS1 and SOCS3 also
have the ability to directly inhibit JAK1, JAK2, and TYK2 (but not JAK3) by blocking the
substrate-binding groove on JAK, thereby acting as a pseudosubstrate. SOCS1 is the most
influential SOCS family member, and is the principal regulator of numerous cytokines involved in
the immune response, in particular IFN-γ (Liau et al. 2018). Genetic deletion of SOCS1 is fatal, and
its down-regulation plays a role in the progression of cancer. Each SOCS protein is also specific for
only a subset of cytokines; for example, SOCS3 is prompted by IL-2, -3, -4, -6, -7, -9, -10, -11, -12,
-13, -21, and -22, as well as granulocyte-colony stimulating factor (G-CSF), granulocyte-macro­
phage colony-stimulating factor (GM-CSF), leukemia inhibitory factor (LIF), PRL, IFN-α, IFN-γ,
GH, EPO, thyroperoxidase, oncostatin M (OSM), calcitonin-1 (CT1), ciliary neurotrophic factor
(CNTF), and leptin (Morris, Kershaw, and Babon 2018).
There are also members of the PIAS family, which are able to inhibit the JAK-STAT pathway
by binding to and inhibiting STAT dimers, preventing DNA binding and gene transcription.
However, the exact mechanism by which PIAS proteins do their regulative functions remains
unrevealed. There are several PIAS family members, including PIAS1, PIAS3 (KchAP), PIASy,
and PIASx (ARIP3) (O’Shea and Plenge 2012). PIAS family members possess several domains,
including a serine/threonine rich domain located at the C-terminus (responsible for target binding),
a Zn-binding RING–finger-like domain (RLD, responsible for SUMO transfer), and a conserved-
SAP domain (scaffold attachment factor A/B, Acinus, PIAS) near the N-terminus (important part for
target binding via scaffold/matrix attachment regions, S/MARs) (Seif et al. 2017).

1.7 Failure of JAK Regulation Leads to Disease


Deregulation of cytokines and/or their downstream signaling pathways are at the root of many
human disorders and diseases, including asthma, severe combined immunodeficiency (SCID), and
various cancers (Gorby et al. 2018; Hammaren et al. 2019). A number of these disorders are because
of point mutations, as in the case of SCID (mutation and loos in function of JAK3). Oncogenic JAK2
rearrangements to multiple fusion gene partners have been identified in acute lymphoblastic leukemia
(ALL), atypical chronic myeloid leukemia (CML), acute myeloid leukemia (AML), myeloprolifera­
tive neoplasms (MPN), and/or Hodgkin lymphoma (Hammaren et al. 2019).
A number of gain-of-function mutations that have been identified in JAK1, JAK2, and JAK3
are responsible for approximately 20% of T-ALL cases and to a lesser extent in B-ALL or
hepatocellular adenoma (Hammaren et al. 2019). In some cases, mutations in other components
and JAK-STAT regulatory molecules can be the cause of the disease, as is the case in 1%–5% of
primary myelofibrosis (PMF) and essential thrombocythemia (ET). In these cases, there is a point
mutation in the thrombopoietin receptor (TPOR) gene MPL, which enables ligand-independent
activation of JAK2 signaling (Leroy et al. 2016). Frameshift mutations are also frequent,
accounting for approximately 13% of MPN cases, causing thrombopoietin-independent TPOR–
JAK2 signaling. In this instance, a frameshift mutation in calreticulin (CALR), an endoplasmic
reticulum protein, results in a highly charged protein capable of binding to TPOR, thus activating
the receptor (Hammaren et al. 2019). Apart from mutations resulting in JAK-STAT hyperactivity,
JAK-STAT loss-of function mutations have also been described. Loss-of-function mutations in
JAK3 or JAK3-associated receptors IL-7R and common gamma chain (IL-2Rγ) cause autosomal
recessive SCID. TYK2 deficiency has been reported in primary immunodeficiency (PID) (Hammaren
et al. 2019).

1.8 Conclusion
There is a shared interaction between external actions and internal reactions that enables a cell to
remain viable and survive. Binding of extracellular cytokines and growth factors to their receptor
8 JAK-STAT Signaling in Diseases

sets off a cascade of responses and activates cellular signaling pathways, leading to gene transcription
and ultimately cellular proliferation, differentiation, activation/inhibition, and survival/apoptosis. The
JAK-STAT pathway plays a fundamental role in the transfer of extracellular signals from membrane
receptors to the nucleus. Over fifty cytokines make use of the JAK-STAT pathway by binding to type
I or type II receptors to carry out their effects. Several regulators modulate the function of the JAK­
STAT pathway, such as protein tyrosine phosphatases, SOCS, and PIAS family members. There is still
a lack of knowledge concerning the complete molecular understanding of JAK activation and
inhibition. Any dysregulation in the JAK-STAT pathway may lead to various pathologies. Better
knowledge of these mechanisms may be the answer in the development of therapeutic strategies to
target the JAK-STAT pathway to treat various immunological disorders and cancers.

REFERENCES
Ferrao, R., H. J. Wallweber, H. Ho, et al. 2016. The structural basis for class II cytokine receptor
recognition by JAK1. Structure 24 (6):897–905.
Filtz, T. M., W. K. Vogel, and M. Leid. 2014. Regulation of transcription factor activity by interconnected
post-translational modifications. Trends Pharmacol Sci 35 (2):76–85.
Gorby, C., J. Martinez-Fabregas, S. Wilmes, and I. Moraga. 2018. Mapping determinants of cytokine
signaling via protein engineering. Front Immunol 9:2143.
Hammaren, H. M., A. T. Virtanen, J. Raivola, and O. Silvennoinen. 2019. The regulation of JAKs in
cytokine signaling and its breakdown in disease. Cytokine 118:48–63.
Hatakeyama, S. 2012. Ubiquitin-mediated regulation of JAK-STAT signaling in embryonic stem cells.
JAKSTAT 1 (3):168–75.
Hendry, L., and S. John. 2004. Regulation of STAT signalling by proteolytic processing. Eur J Biochem
271 (23–24):4613–20.
Jatiani, S. S., S. J. Baker, L. R. Silverman, and E. P. Reddy. 2010. Jak/STAT pathways in cytokine signaling
and myeloproliferative disorders: approaches for targeted therapies. Genes Cancer 1 (10):979–93.
Leroy, E., J. P. Defour, T. Sato, et al. 2016. His499 regulates dimerization and prevents oncogenic
activation by asparagine mutations of the human thrombopoietin receptor. J Biol Chem 291
(6):2974–87.
Liau, N. P. D., A. Laktyushin, I. S. Lucet, et al. 2018. The molecular basis of JAK/STAT inhibition by
SOCS1. Nat Commun 9 (1):1558.
Lim, Cheh Peng, and Xinmin Cao. 2006. Structure, function, and regulation of STAT proteins. Mol
Biosyst 2 (11):536–50.
Morris, R., N. J. Kershaw, and J. J. Babon. 2018. The molecular details of cytokine signaling via the JAK/
STAT pathway. Protein Sci 27 (12):1984–2009.
Murray, P. J. 2007. The JAK-STAT signaling pathway: input and output integration. J Immunol 178
(5):2623–9.
O’Shea, J. J., and R. Plenge. 2012. JAK and STAT signaling molecules in immunoregulation and immune-
mediated disease. Immunity 36 (4):542–50.
Schindler, C., D. E. Levy, and T. Decker. 2007. JAK-STAT signaling: from interferons to cytokines. J Biol
Chem 282 (28):20059–63.
Seif, F., M. Khoshmirsafa, H. Aazami, M. Mohsenzadegan, G. Sedighi, and M. Bahar. 2017. The role of
JAK-STAT signaling pathway and its regulators in the fate of T helper cells. Cell Commun Signal
15 (1):23.
Wilks, A. F., and A. G. Harpur. 1994. Cytokine signal transduction and the JAK family of protein
tyrosine kinases. Bioessays 16 (5):313–20.
2
The Structure-Function Bonhomie of
JAK-STAT Molecules

Ritobrata Goswami
School of Bioscience
Indian Institute of Technology Kharagpur
Kharagpur, India

2.1 Introduction
The existence of life requires the ability to comprehend and respond to external signals. These
signals are intercepted by numerous cell-surface receptors that utilize a range of intracellular
signaling pathways to communicate with the nucleus at the cellular level. The culmination of these
signaling ensue an appropriate response mediated by peripheral sensory organs via the central
nervous system. The Janus kinase (JAK)-signal transducers and activators of transcription
(STAT) pathway used by plants and animals, including flies, is one of the few signal transduction
pathways that transduce a multitude of signals required for development and homeostasis
(Rawlings, Rosler, and Harrison 2004). Activation of JAK induces various physiological events
including, but not limited to, cell proliferation, migration, differentiation, and apoptosis. JAK­
STAT pathway’s association is evident in 12 major cancers (Vogelstein et al. 2013). JAK-STAT
pathway is one of the major signaling pathways employed by multiple growth factors and
cytokines (Liongue and Ward 2013). These cellular events are critical to hematopoiesis, develop­
ment of immune system, and subsequent functions. JAK-STAT signaling is dependent upon
tyrosine phosphorylation and harbors comparatively simple signal transduction pathway requir­
ing few components. Activation of JAK-STAT pathway ensues when growth factors or cytokines
bind to their cognate receptors leading to multimerization of receptor subunits (Rawlings, Rosler,
and Harrison 2004). The multimerization of receptor subunits can form homodimers that lead to
close proximity of the receptor-associated JAK molecules. Via transphosphorylation, JAKs
phosphorylate each other on tyrosine residues, a cue for their activation. Activated JAKs then
phosphorylate on tyrosine residues of cytokine or growth-factor receptors paving the way for
binding of proteins bearing SH2 domains. With the help of SH2 domains, STAT molecules bind
to the phosphorylated tyrosine residues on the cognate receptor. STAT molecules dissociate from
the receptor after being tyrosine-phosphorylated by JAKs. Following activation, STAT molecules
form either homo or hetero-dimers and SH2 domain of each STAT molecule binds to the
phosphotyrosine residue of its corresponding STAT. Subsequently, the STAT dimer translocate
to the nucleus, bind to the genes bearing consensus STAT-binding sequence, and activate gene
transcription (Figure 2.1). Thus, the JAK-STAT signaling cascade is a direct mechanism to
translate an extracellular signal into a transcriptional response. The canonical JAK-STAT
proteins are inactivated by negative regulators, including SH2-containing protein tyrosine phos­
phatase (SHP) and suppressor of cytokine signaling (SOCS) proteins (Liongue et al. 2012). While
phosphorylation can be denoted as an “on” switch (binary 1), dephosphorylation can be
attributed as an “off” switch (binary 0). In this chapter, the structure–function relationship of
JAK-STAT signaling has been discussed.

9
10 JAK-STAT Signaling in Diseases

FIGURE 2.1 Brief overview of JAK-STAT domain structure. The JAK-STAT signaling pathway has been shown in
between the domain structures JAK (shown on top) and STAT (shown at the bottom).

2.2 The Beginning


The journey of JAK-STAT pathway started with the discovery of “interferons” way back in 1957
by Issacs and Lindenmann (Isaacs and Lindenmann 1957). As the moniker suggests, interferons
(IFNs) “interfered” with the replication process of virus. Three decades later, studies involving the
induction of mRNA and translation of proteins by interferons staged the set for identifying
members of JAK-STAT signaling pathway. JAK family has four members: JAK1, JAK2, JAK3,
and TYK2. STAT family has seven members: STAT1, STAT2, STAT3, STAT4, STAT5A,
STAT5B, and STAT6. During 1984–1988, IFN-α-induced transcriptional stimulation of specific
genes was characterized and IFN-dependent promoters were identified (Stark and Darnell 2012).
The first members of the JAK family were identified in 1989, while cDNA clones for the first
STAT family member was discovered in 1992 (Wilks 1989, Shuai et al. 1992). The Janus of Janus
kinase comes from two-faced Roman God of doorways, “Janus”, since JAK possesses two near-
identical phosphate-transferring domains (Wilks 1989). During 1989–1991, JAK1, JAK2, and
TYK2 were identified, while JAK3 sequencing was completed in 1994 (Stark and Darnell 2012).
Sequencing of STAT3, 4, 5A, 5B, and 6 were completed in the same year (Larner et al. 1993,
Ruff-Jamison, Chen, and Cohen 1993, Sadowski et al. 1993, Silvennoinen et al. 1993, Akira et al.
1994, Zhong, Wen, and Darnell 1994a, 1994b, Jacobson et al. 1995). The functional and
structural domains of STATs were described during 1995–1998 (Hoey and Schindler 1998).
STATs were first crystalized in 1998 followed by identification of target genes (Becker, Groner,
and Muller 1998). The onset of twenty-first century was when the underlying importance of the
signaling pathway was linked to several disorders that were associated with JAK-STAT mutations
and gained momentum.
JAK-STAT Structure Function 11

2.3 Structure and Activation of JAK


Cytokine or growth-factor receptors lack intrinsic kinase domain, thus depending on JAK for
transmitting signal to the cytoplasmic components. Structurally, JAKs have seven domains known
as JAK homology regions (JH1-7) (Bousoik and Montazeri Aliabadi 2018) (Figure 2.1). Func­
tionally, JAKs have been shown to possess four domains: catalytically active tyrosine kinase
domain, the pseudotyrosine kinase domain, the Src homology 2 (SH2) domain, and the FERM
domain (Bousoik and Montazeri Aliabadi 2018). JH1 encodes the kinase domain, while JH2 has
a pseudo-kinase domain that controls the kinase activity of JH1 (Manning et al. 2002, LaFave
and Levine 2012). The JH2 domain activity has been observed specifically for JAK2 and has been
suggested to possess auto-inhibitory function (Ungureanu et al. 2011). The SH2 domain is
encoded by JH3 and JH4 regions, while JH5-7 is responsible for receptor binding (Cornejo,
Boggon, and Mercher 2009). The JH3-7 regions form the N-terminal of JAKs, while JH1-2 forms
the C-terminus of the domain (Yamaoka et al. 2004). JAK activation occurs upon ligand-
mediated receptor multimerization because two JAKs are brought into close proximity, allowing
trans-phosphorylation. Binding of the ligand to cytokine receptor revamps the receptor–JAK
dimers, which brings the JAKs close enough to transphosphorylate the partner JAK in the dimer
at JH1. The activated JAKs subsequently phosphorylate additional targets, including both the
receptors and the major substrates, STATs (Yamaoka et al. 2004). The specificity of ligand-
mediated STAT activation is dependent on the interactions between SH2 domain and phospho­
tyrosine residues. The current understanding of JAK domains will be elaborated in the subsequent
sections.

2.4 Structure of STAT and Translocation to the Nucleus


STATs function both as signal transducers and transcription factors. STATs are latent transcrip­
tion factors that reside in the cytoplasm until activated. Like JAKs, STATs also possess distinct
conserved domains: the N-terminal coiled-coil domain, DNA-binding domain, a linker, SH2
domain, and C-terminal transactivation domain (Figures 2.1 and 2.2). The transactivation
domain interacts with multiple partners forming transcriptional complexes. Tyrosine at position
705 is important when it comes to STAT activation (Yu and Jove 2004). Other than STAT2 and
STAT6, a conserved serine at position 727 is another site of phosphorylation that controls STAT
transcriptional activity (Decker and Kovarik 2000). The DNA-binding domain is the central
region that controls DNA-binding selectivity for each STAT molecule. Between position 600 and
700 amino acid residues, the SH2 domain is located that aids the dimer formation between two
activated STAT molecules via SH2-phosphotyrosine interaction (Horvath, Wen, and Darnell
1995). The linker region is located between the amino acid residues 500 and 575 (Subramaniam
et al. 2013).
After activation by tyrosine phosphorylation, STATs become dimerized and translocate into the
nucleus, where they act as transcription factors. While most STATs form homodimers, evidences
support heterodimer formation of STAT1/2, STAT1/3, and STAT5A/B (Delgoffe and Vignali
2013). STAT1 has been reported to exist as pre-formed homodimers, and phosphorylation
induces reorientation. STAT molecules face restricted movement while translocating from cyto­
plasm to nucleus by the nuclear envelope owing to molecular weight >40 kDa (Mertens et al.
2006). Thus, STATs require facilitated transport mediated by specific receptor molecules that are
known as importins (adaptor proteins) bearing α and β subunits. Using nuclear localization
sequence (NLS), STATs bind to importin α (Gorlich and Mattaj 1996, Jans, Xiao, and Lam
2000). Following interaction of both the importin subunits, STAT–importin complex gets docked
at the nuclear pore complex (NPC) (Gorlich and Mattaj 1996). The NPCs form high-order
octagonal channels that are composed of proteins called nucleoporins, some of which contain
hydrophobic phenylalanine/glycine (FG)-rich repeat motifs (Wolf and Mofrad 2008). The energy
12 JAK-STAT Signaling in Diseases

FIGURE 2.2 STAT domain structure. Minor differences in various human STAT molecules has been depicted.

for the translocation is provided by Ran, a GTPase. The intracellular RanGTP/RanGDP gradient
essentially drives this active transport, which allows the accumulation of cargo against
a concentration gradient. In the cytosol, the concentration of the GTP form of Ran is low due
to nucleotide hydrolysis by RanGAP, which is cytoplasmically localized by RanGTPase-activating
protein (Reich 2013). A differential requirement of importin α has been attributed for STAT
translocation (Goldfarb et al. 2004). Out of the six human importin α molecules, STAT1 requires
importin α5, while STAT5 requires importin α3 (Sekimoto et al. 1997, Liu, McBride, and Reich
2005). One study even argued that STATs may not be having functional NLS (Subramaniam,
Torres, and Johnson 2001). Furthermore, it has also been suggested that NLS-binding site on
STAT1 and STAT3 differs from its binding site in other proteins. STAT3, 5, and 6 could also be
translocated to the nucleus in unphosphorylated form (Reich 2013). Carrier-dependent export of
unphosphorylated STATs requires the specific export receptor, CRM1 (Fornerod et al. 1997).
Translocation of unphosphorylated STATs requires direct interaction with nucleoporins and is
carrier and energy-free, devoid of any help from importins. This process also does not require any
activation by cytokines.

2.5 JAK Domains


As mentioned previously, JAKs contain specific domains. In this section, we discuss about the
domains found in JAKs.

2.5.1 Protein Tyrosine Kinase Domain


The JH1 domain is the PTK domain of the JAK protein. Similar to other protein tyrosine
kinases, JAKs possess an “activation loop” that regulates kinase activity, which is a major site of
autophosphorylation (Toms et al. 2013). Using crystal structures of JAK2 and JAK3, it has been
JAK-STAT Structure Function 13

demonstrated that the PTK has an open conformation representing the active conformation of
the protein. JAK2 and JAK3 crystal structures have been solved at 2A° and 2.5A°, respectively
(Boggon et al. 2005, Lucet et al. 2006). JAK2 PTK domain has a small N-terminal lobe,
comprised of one α-helix and five anti-parallel β strands, and large C-terminal lobe comprised of
eight α-helices (Taylor et al. 1992). The JAK2 kinase domain is relatively an “open” conforma­
tion, but the ATP-binding site is relatively “closed” when compared to other kinases (Wilks 2008).
The two lobes of the JAK2 kinase domain are slightly twisted with respect to other kinases. The
ATP-binding site is relatively less accessible as a consequence. A loop structure located between
amino acids 1056–1078, termed the JAK2 kinase insertion loop, is rather a unique feature not
observed in any other kinases (Wilks 2008). JAK3 has ~60% sequence identity with JAK2 and
most of the amino acid residues are conserved in the ATP-binding cleft (Boggon et al. 2005).

2.5.2 Pseudokinase Domain


JAKs are named after Roman God “Janus” as they have a kinase and “kinase-like” pseudokinase
domain. The pseudokinase domain has been suggested to have critical positive and negative
regulatory roles (Saharinen, Takaluoma, and Silvennoinen 2000, Saharinen, Vihinen, and Silvennoi­
nen 2003). Despite lacking key residues that are shown to be important for catalytic activity, this
domain has retained the capability to bind nucleotides in the presence of divalent cations and
mediates essential functions in regulating catalytic activity. The pseudokinase domain shares high
degree of sequence similarity with the kinase domain, but residues required for phosphotransferase
activity are modified from the canonical motifs. Among the four JAK molecules, only the JAK2
pseudokinase domain possesses weak catalytic activity and autophosphorylates itself in cis on two
auto-inhibitory phosphorylation sites, Ser523 and Tyr570 (Argetsinger et al. 2004, Ishida-Takahashi
et al. 2006, Ungureanu et al. 2011). Evidence of JAK2-specific mechanism of activation comes from
the fact that those residues are not conserved among the other JAKs. Two models of JAK activation
have been hypothesized. The pseudokinase domain binds and inhibits the kinase domain either in cis
or in trans (Babon et al. 2014). The JAK pseudokinase domains act as protein interaction modules,
whose primary function is to bind and inhibit the JH1 domain activation. However, many studies
have reported conflicting findings on the mechanism of inhibition. Furthermore, it is suggested that
nucleotide binding to these pseudokinase domains primarily function to modulate the overall
conformation of the JAK (Hammaren et al. 2015). The amino acid residues forming the “activation
loop” in the JAK2, JAK1, and TYK2 pseudokinase domain form a well-defined conformation that
resembles inactive protein kinase structures. These structures have been hypothesized to block the
putative substrate binding site, subverting the fact that a rearrangement of JAK2 pseudokinase
activation loop is required for phosphoryl-transfer activity. The importance to arrange a specific
conformation to allow intermolecular interactions is evident from crucial stabilization mediated by
ATP binding on purified recombinant JAK1 and JAK2 pseudokinase domains (Hammaren et al.
2015). In polycythemia vera (PV) patients, mutations in the JAK2 pseudokinase-SH2 linker region,
including K539I, were recently characterized that reportedly led to constitutively activated JAK2
(Zhao et al. 2009). Two distinct mutational hotspots within the JAK2 pseudokinase domain have
been identified based on functional studies of activation (Babon et al. 2014). Close to V617F,
mutations point to the N-terminal lobe of the domain. The pseudokinase-SH2 linker region was
found to confer factor-independent growth. However, there was no detectable increase in the
catalytic activity (Babon et al. 2014). Interestingly, the crystal structure of the JAK2 pseudoki­
nase domain possessing the activating mutant, V617F, indicates toward increased structural
integrity within this domain leading to pathogenesis (Babon et al. 2014). In contrast, another
structure of the JAK1 pseudokinase domain demonstrated that the αC region length can be
variable regardless of the presence of the subsequent activating mutation, V658F (Toms et al.
2013). Examples exist to support JAK phosphorylation on other sites that appear to be
important for regulation of catalytic activity. For example, Y785 in JAK3 and Y813 in JAK2
denote other important sites of autophosphorylation (Kurzer et al. 2004). These sites recruit
SH2-domain-containing adapter molecules such as SH2-Bb. For JAK2, this could be a mechanism
14 JAK-STAT Signaling in Diseases

to enhance catalytic activity. APS, another related adapter, also binds JAK2, but it negatively
regulates the activity in contrast (O’Brien, O’Shea, and Carter-Su 2002).

2.5.3 SH2 Domain


A sequence-specific phosphotyrosine-binding module, the SH2 domain is present in many signaling
molecules. In humans, SH2 domains constitute the largest class of phosphotyrosine-specific recogni­
tion domains (Kaneko et al. 2012). The SH2 domain controls cellular localization, substrate
recruitment, and regulation of kinase activity in proximity to the SH1 domain located at the
N-terminal of cytoplasmic tyrosine kinases (Filippakopoulos, Muller, and Knap 2009). The position­
ing of an N-terminal SH2 domain followed by a kinase domain is strongly conserved in all family
members of cytoplasmic tyrosine kinases (Filippakopoulos, Muller, and Knap 2009). Evolutionary,
this might have resulted in an invariant signaling unit associated with the occurrence of tyrosine
phosphorylation. In eukaryotic cells, protein kinase activity is tightly regulated. Majority of the
kinases are kept in an inactive state that can be rapidly activated by interaction with regulatory
elements such as interacting with proteins or domains located outside the catalytic domain to ensure
specific propagation of cellular signals (Wagner et al. 2013). It is suggested that this specific domain
arrangement are meant to target kinases to their substrates. Initial structural studies indicated that
SH2 domain stabilized the inactive state of Src family members (Filippakopoulos, Muller, and Knap
2009). However, subsequent biochemical characterization studies implied that the presence of the
SH2 domain is required for catalytic activity. This event stabilizes the active state of many non-
receptor tyrosine kinases. Furthermore, signaling selectivity is rapidly increased by recruiting
kinases to signaling complexes (Ardito et al. 2017). Interactions with secondary substrate
localization sites aid in selective targeting of substrates. Post-translational modifications also
help in quick propagation of signals. The SH2 domain has additional regulatory functions
including allosteric regulation of the kinase catalytic domain that is commensurate with
enhanced signaling complexities in higher order eukaryotes (Filippakopoulos, Muller, and
Knap 2009). Using a large number of biophysical and biochemical studies, the selectivity of
SH2 domains for their phosphotyrosine substrates have been determined. Approximately 50% of
the binding affinity comes from the phosphate moiety of the phosphotyrosine residue (Kaneko
et al. 2012). Directed phosphopeptide library-screening based studies indicated that residues in
positions from −2 to +4, relative to the phosphotyrosine, regulate binding specificity (Bradshaw,
Mitaxov, and Waksman 1999, Machida et al. 2007). In contrast, studies involving some co-
crystal structures suggested larger contact interfaces spanning from residues −6 to +6 (Liu et al.
2006). The overall structure of the JAK-family SH2 domain is comprised of a central β-sheet
flanked by two α-helices that resembles a canonical phosphotyrosine-binding SH2 domain
(Ferrao and Lupardus 2017). In canonical SH2 domains, two loops that go around the SH2-αB
helix form a hydrophobic groove at positions +3 and +5 relative to the pTyr opening up the
binding site for specificity determining residues (Bradshaw and Waksman 2002). In addition, the
conserved, phosphate-coordinating arginine residue located at the base of the pTyr binding
pocket is conserved in all JAKs, except TYK2, where it has been replaced with a histidine residue
(Ferrao and Lupardus 2017). In non-receptor tyrosine kinases, the conserved SH2–kinase unit is
flanked by additional domains that could include the N-terminal flanking region, a Src homol­
ogy 3 (SH3) domain, a second SH2 domain, a sequence of PH (plecstrin homology)–BTK–SH3
domains, and the FERM domain (Filippakopoulos, Muller, and Knap 2009). We discuss the
FERM domain in the next section.

2.5.4 FERM Domain


The FERM domain is conserved as a number of FERM domain-containing proteins, which are
found in present-day plants and basal eukaryotes. The FERM domain is named after
the founding protein members identified with this domain (band 4.1, ezrin, radixin, and moesin)
(Ferrao and Lupardus 2017). This domain is comprised of three subdomains: F1, F2, and F3,
JAK-STAT Structure Function 15

which are structurally similar to ubiquitin, acyl-CoA binding, and plecstrin homology­
phosphotyrosine-binding domains, respectively (Tepass 2009). The FERM domain mediates
protein–protein interactions that include adaptor and scaffolding interactions with membrane-
bound proteins. JAK-STAT signaling components also consist of several other domains that
provide accessory functions. The SOCS box is approximately 40-residue motif that mediates
interactions with proteasomal degradation pathway components, thus regulating protein half-life
(Kile et al. 2002). FERM-domain containing proteins can be divided into three broad groups:
(1) talins and kinesins; (2) ERMS, GEF, kinases, and phosphatases; and (3) myosins and KIRT
(Frame et al. 2010).
Structural model of the JAK FERM and SH2 domains and the mechanism of JAK interac­
tion with the receptors remained a challenge till five years back, owing to poor expression and
solubility of JAK protein crystallization (Lupardus et al. 2011). In 2014, co-purification with
a stabilizing receptor led to an initial X-ray crystal structure of the human TYK2 FERM and
SH2 domains, followed by crystal structures of the human JAK1 and JAK2 FERM–SH2
fragments in 2016 (Wallweber et al. 2014, Ferrao et al. 2016, McNally, Toms, and Eck 2016,
Zhang, Wlodawer, and Lubkowski 2016). These structures revealed that the JAK FERM and
SH2 domains are tightly associated to form a single receptor-binding module (Ferrao and
Lupardus 2017). Molecular modeling suggested that this domain could adopt a classical
FERM-domain structure. Two additional loops are found within the characteristic FERM
structure that are absent from other FERM domains, which are not well conserved between
JAK family members. The overall structural organization of the JAK FERM is similar to
canonical FERM domains. These subdomains pack into a canonical trilobed FERM architec­
ture (Ferrao and Lupardus 2017).
While the overall domain topology is conserved, JAK FERMs have several key differences
compared to FERM domains from the ezrin/radixin/moesin and FAK families that lead to close
contact and interaction with the SH2 domain (Ferrao and Lupardus 2017). L1, the elongated
linker joins the F1 and F2 domains to construct a major portion of the interaction surface
between the FERM and SH2 domains (Ferrao and Lupardus 2017). L1 is extended between 29
and 42 residues in JAKs that is different from the typical 13 and 15 residues in classical FERMs
(Ferrao and Lupardus 2017). Apart from the L1 linker, the SH2 domain also stacks against the
F1-α1 helix, L2 (the highly conserved F3–SH2 linker), and the L3 linker (SH2–pKD linker)
(Wallweber et al. 2014). The difference between JAK FERM and classical FERM is also evident
in the F2-subdomain structure (Ferrao and Lupardus 2017). The extension of the N-terminal
end of F2-α2 helix by one turn in JAK2 and two turns in TYK2 and JAK1 occurs due to
additional residues in the linker region between F2-α1 and F2-α2 (Ferrao and Lupardus 2017).
In JAK FERMs, the C-terminal end of the F2-α2 and the N-terminal end of F2-α3 helices also
increased by one turn. Multiple loops within the F3 subdomain display significant variability
between different JAK isoforms (Wallweber et al. 2014). The first loop contains 12 amino acids
in JAK2 between F3-β1 and F3-β2 strands that are the mostly disordered in the crystal structure
(Ferrao and Lupardus 2017). In JAK1 and TYK2, the F3-β1/β2 loop contains 22 amino acids
and 35 amino acids, respectively (Ferrao and Lupardus 2017). F3-β1/β2 loop of JAK1 forms
a β hairpin that packs against F3-β7, extending the β-sheet by a single strand, F3-β1 (Ferrao
and Lupardus 2017). Similarly in TYK2, the loop also forms a single-strand packing against
F3-β7, with an additional visible loop. In TYK2, the C-terminal end of the linker is unstruc­
tured (Ferrao and Lupardus 2017). Another difference lies between the JAK FERM and
classical FERM in the large insertions at this position that are specific to JAK-family FERM
domains (Ferrao and Lupardus 2017). The classical FERM domains contain only a short loop
at this position. An additional disordered loop of variable length is located in between F3-β3
and F3-β4. The linker lengths of F3-β3/β4 region are 34, 12, and 44 amino acids in JAK1,
JAK2, and TYK2, respectively (Ferrao and Lupardus 2017). These insertions are also not
present in canonical FERM domains thereby differing from the JAK FERM domain. However,
the length and sequence identities of these JAK F3 insertions are tightly conserved between all
higher eukaryotic species.
16 JAK-STAT Signaling in Diseases

2.6 STAT Domains


As described previously, STAT proteins have a modular structure with six well-defined domains
(Baker, Rane, and Reddy 2007). The functions of STATs are mediated by N-terminal region since
small deletions in this region do away with the STAT phosphorylation (Baker, Rane, and Reddy
2007). The N-terminal region of STAT is well conserved among family members and cooperates
with the DNA-binding domain (Baker, Rane, and Reddy 2007). Additionally, this domain also
has major roles in nuclear import, export, and receptor binding. Coiled-coil domain of STAT
interacts with regulatory proteins, adopting α-helical conformation in doing so (Kisseleva et al.
2002). This domain functions in receptor binding. Similar to the N-terminal domain, the DNA-
binding domain is also highly conserved among STAT family members (Baker, Rane, and Reddy
2007). Apart from STAT2, all STAT homodimers differentially bind more than 10 related
γ-activated sequence elements characterized by the consensus sequence, TTNCNNNAA
(Horvath, Wen, and Darnell 1995, Xu, Sun, and Hoey 1996). The linker domain functions as
a spacer to retain true conformation between the dimerization and DNA-binding domains (Baker,
Rane, and Reddy 2007). The SH2 domain, which is also found in JAKs, is highly conserved
among the STAT molecules. In playing a key role in STAT signaling, the SH2 domain recruits
STATs to activated receptor complexes. Moreover, the SH2 domain is also meant for interaction
with JAK and Src kinases (Baker, Rane, and Reddy 2007). The same domain is also required for
dimerization (both homo- and heterodimerization) of STAT molecules. This event leads to nuclear
localization and DNA-binding activities (Baker, Rane, and Reddy 2007). The transactivation
domain unsurprisingly modulates the transcriptional activation of target genes but varies among
family members. Modulation of transcriptional activation is evident when the C-terminally
truncated isoforms of STAT3, 4, and 5 minus the portions of their transactivation domains, act
as dominant-negatives (Schindler and Strehlow 2000).
In an unstimulated cell, STATs are monomeric and remain in unphosphorylated state. Cytokine
stimulation induces phosphorylation of tyrosine residues on the receptor that serve as docking
sites for STATs via their SH2 domains. Once bound to the receptor, all members of the STAT
family become tyrosine phosphorylated in response to cytokine stimulation at a conserved
C-terminal tyrosine residue. For example, phosphorylation of Y694 in STAT5 can be mediated
by growth-factor receptors as well as by JAK and Src kinases, depending on the nature of the cell
type and the ligand/receptor interactions (Nosaka et al. 1999). This phosphorylation event induces
STAT homo- and heterodimerization via the interaction of the SH2 domain of one STAT molecule
with the phosphotyrosine residue of another. Once phosphorylated, the dimerized STATs are then
able to translocate to the nucleus. In addition to tyrosine phosphorylation, several STATs are
regulated by serine phosphorylation at a conserved PSMP motif, which is located in the transacti­
vation domain (Jatiani et al. 2010). C-terminal serine phosphorylation is stimulated by several
cytokines and is mediated by serine/threonine kinases that include extracellular responsive kinase,
p38, JNK, and protein kinase C-δ (Khwaja 2006). This phosphorylation event positively regulates
the transactivation potential of these proteins. One study has revealed that the specificity of
signaling by STAT1 depends on SH2 and C-terminal domains that regulate Ser727 phosphorylation
(Kovarik et al. 2001). This culminates in differential regulation of specific target gene expression. To
acquire complete activation of STAT1, phosphorylation needs to happen at both Y701 and S727
residues (Kovarik et al. 2001). Interestingly, S727 phosphorylation of STAT1 in IFN-γ-treated
mouse fibroblasts requires an intact SH2 domain and phosphorylation of Y701 without any
requirement for p38 MAPK, ERK-1 and -2, and JNK (Kovarik et al. 2001). In contrast,
UV-induced STAT1 phosphorylation on S727 does not need SH2 domain–phosphotyrosine interac­
tions, but require p38 MAPK (Kovarik et al. 2001). Mutation of S727 differentially affected IFN-γ
target genes, at the level of both basal and induced expression. Furthermore, S727 of STAT3 is
phosphorylated in response to stimuli that differ from those for STAT1 S727, and transfer of the
STAT3 C terminus to STAT1 changed the pathway specificity of STAT1 Ser727 phosphorylation to
that of STAT3 (Kovarik et al. 2001). This series of experiments show that STAT C terminus
JAK-STAT Structure Function 17

contributes to the specificity of cellular responses by linking individual STATs to different kinase
pathways and doing so through an intrinsically different requirement for serine phosphorylation at
different target gene promoters.

2.7 Negative Regulation of JAK-STAT Pathway


It is evident that cytokines or growth factors signal through JAK-STAT pathway. Unsurprisingly,
multiple studies over several years have shown that dysregulated cytokine receptor signaling is the
foundation of many mammalian diseases. Normal cytokine signaling is dependent on the complex
tuning of cytokine-dependent signals. There are mechanisms that work against these signals to
allow a cell to be attuned to additional extracellular signals. There are multiple mechanisms by
which cytokine signaling via JAK-STAT pathway is negatively regulated. In this section, we
deliberate upon the negative regulation of JAKs via phosphatase, inhibition of STATs by protein
inhibitors of activated STATs (PIAS) and suppressor of cytokine signaling (SOCS) proteins.

2.7.1 Effect of Protein Tyrosine Phosphatases (PTPS)


Tyrosine phosphorylation leads to the signaling that emanates ensuing cytokine–cytokine receptor
interaction. The event of JAKs phosphorylating each other on tyrosine residues in activation loops is
known as transphosphorylation. Due to the critical importance of JAK tyrosine phosphorylation for
signal transduction, dephosphorylation of tyrosine residues is a mechanism employed by cells to
attenuate kinase activity of JAKs. Several protein phosphatases act as modulators of JAK-STAT
signaling. They include the SH2 domain-containing tyrosine phosphatases (SHPs), protein tyrosine
phosphatase 1B (PTP1B), T-cell protein tyrosine phosphatase (TC-PTP), and protein tyrosine phospha­
tase-basophil like (PTP-BL). Among these, SHPs are well characterized. The classical PTPs are
Cys-based class I PTPs that are true tyrosine-specific PTPs (Alonso et al. 2004). Based on their cellular
localization, the classical PTPs are divided into transmembrane (TM) and non-transmembrane (non-
TM) families (Tonks 2006). Few TM PTPs have been identified in T cells. Non-TM PTPs are better
characterized than their TM counterparts (Lorenz 2009). Non-TM PTPs contain a single catalytic PTP
domain flanked by N-/C-terminal extensions required for localization (Lorenz 2009).
SHPs were the first phosphatases to be identified. The human genome project has identified
more than 100 putative PTPs (Alonso et al. 2004). The major SHPs are SHP-1 and SHP-2. SHP-1
is primarily expressed in hematopoietic cells, while SHP-2 is ubiquitously expressed (Wu et al.
2003). A spliced variant of SHP-1 has been identified as SHP-1L (Jin, Yu, and Burakoff 1999).
Both SHP-1 and SHP-2 are composed of a central catalytic domain, two N-terminal SH2
domains, and a C-terminus (Poole and Jones 2005). The central catalytic domain contains the
typical PTP signature motif VHCSAGIGRTG (Poole and Jones 2005). These phosphatases are
named due to the two SH2 domains located N-terminal to the tyrosine phosphatase domain. The
SH2 domains aid in SHP recruitment to tyrosine-phosphorylated intracellular portions of many
receptors that aid in localization and activity regulation (Pao et al. 2007). SHPs inhibit JAK­
STAT signaling in two ways: by impeding interactions with other effectors and by activating the
elimination of the phosphotyrosine motifs acting as docking sites for effectors (Seif et al. 2017).
The SHP-2 SH2 domains display overlapping binding specificities with the SOCS SH2 domain,
indicating that both negative regulators compete for binding sites on phosphorylated receptors.
The interaction of the SHP SH2 domains with their binding partners is a key step toward the
activation of the phosphatase activity. At the basal state, this activity is repressed as N-terminal
SH2-domain occupies the phosphatase active site of PTP in the intramolecular fashion (Hof et al.
1998). The release of this suppression leads to an activation of the phosphatase upon engagement
of the SH2 domains. SHP-1 is a major negative regulator of receptor signaling in hematopoietic
and epidermal cells (Lorenz 2009). Apart from the SH2 domains, the C-terminus could regulate
the PTP-1 activity of SHP-1. Upon stimulation tyrosine phosphorylation of the C-terminal Y536
18 JAK-STAT Signaling in Diseases

in SHP-1, PTP activity is enhanced (Zhang et al. 2003). Phosphorylation of Tyr564 leads to
a modest increase of SHP-1 activity (Zhang et al. 2003). The role of SHP-2 in growth factor
signaling remains less understood. Additional studies are required to completely elucidate the role
of SHP-2 in hematopoiesis. The intrinsic differences in the catalytic domain allow SHP-2 to portray
different substrate specificities. SHP-2 associates with multiple growth-factor receptors, but few of
those interactions are true SHP-2 substrates. Additionally, SHP-2 could enact as an adapter and has
several phosphotyrosine motifs that can recruit downstream effectors. Phosphorylated tyrosines
have been hypothesized to generate docking sites for SH2-domain containing proteins, which could be
the basis of either an adapter function for SHP-2 or a way for preventing auto-dephosphorylation.
SHP-1L lacks the potential Y564 phosphorylation site, but retains the Y536 site of SHP-1 (Lorenz
2009). SHP-1L, however, has a proline-rich motif in the C-terminus, which could mediate binding to
SH3-domain containing proteins (Lorenz 2009). A similar proline-rich motif is absent in SHP-1, but
can be observed in the C-terminus of SHP-2 (Lorenz 2009). Critical biological functions of SHP-1
and SHP-2 have been demonstrated using mice models.
PTP1B is a negative regulator of the insulin and leptin signaling pathways. PTP1B comprises of
a single catalytic domain with N- or C-terminal extensions (Fischer, Charbonneau, and Tonks
1991). Similar to SHP-1L, PTP1B has a proline-rich region adjacent to the C-terminal and
a hydrophobic region (Liu, Hill, and Chernoff 1996). Substrates are recruited by the interaction of
the proline-rich region with the SH3 domains. The hydrophobic sequence is inserted into the
endoplasmic reticulum membrane and ties up PTP1B to the cytoplasmic side of the ER (Frangioni
et al. 1992). The substrate selectivity of PTP1B is controlled partially by the modular protein
interaction and the spatial restriction imposed by subcellular compartmentalization (Tonks 2003).
PTP1B also acts as a negative regulator of brown fat adipogenesis (Matsuo et al. 2011).
The phosphatases, PTP1B and TC-PTP, share 74% identity between their catalytic domains
(Tonks, Diltz, and Fischer 1988). TC-PTP was characterized almost three decades ago. cDNA of
PTPN2 codes for TC-PTP. TC-PTP, ubiquitously expressed, shows its highest expression in
hematopoietic tissues (Ibarra-Sanchez et al. 2001). TC-PTP could be induced by either a mitogen
(Con A) or an anti-inflammatory cytokine (IL-10) (Rajendrakumar, Radha, and Swarup 1993,
Williams et al. 2004). Expression of TC-PTP results in two isoforms due to alternative splicing. The
two forms share the conservative catalytic PTP domain of 272 amino acids, but differ in the
C-terminal (Kamatkar et al. 1996). The well-characterized form of TC-PTP has 387 amino acids
and has MW of 45-kDa (TC45), while the less-abundant form consists of 415 amino acids with a MW
of 48-kDa and expressed in the ER and the nuclear membrane (Iversen et al. 2002, Bourdeau, Dube,
and Tremblay 2005). The C-terminal of TC45 aids in nuclear localization and also binds to DNA.
However, in general, the C-terminal of TC-PTP negatively regulates the enzyme function. TC-PTP
and PTP1B have complementary specificities for dephosphorylation of JAKs and STATs. Minor
differences between JAK1/JAK3 and JAK2/TYK2 sequences determine the substrate specificity of
PTP1B and TC-PTP. JAK1 and JAK3 activation loops contain either a Thr or Val residue C-terminal
to the tandem phosphotyrosines. In contrast, JAK2 and TYK2 comprise of (E/D)-pY-pY-(R/K)
motif in their activation loops (Myers et al. 2001). This specific motif has been revealed to be the
preferred substrate of PTP1B. Both PTP1B and TC-PTP participate in STAT dephosphorylation
(Bohmer and Friedrich 2014). Biological roles of PTP1B and TC-PTP have been characterized using
mouse knockout studies. Mice deficient in TC-PTP gene exhibit multiple defects in the lymphoid
population (You-Ten et al. 1997). TC-PTP-deficient mice die neonatally from a systemic inflammatory
disease characterized by mononuclear cell infiltration. PTP1B-deficient mice demonstrate hypersensi­
tivity to various growth factors (You-Ten et al. 1997, Hendriks et al. 2008). The other phosphatase, PTP­
BL, dephosphorylates STAT1, STAT3, STAT4, STAT5, and STAT6 both in vitro and in vivo.

2.7.2 Effect of Protein Inhibitor of Activated STATs


PIAS proteins are activation-suppressing proteins for STATs incorporating transcriptional regula­
tion of genes. The PIAS family of proteins consists of five proteins in mammals: PIAS1, PIASxα,
JAK-STAT Structure Function 19

PIASxβ, PIAS3, and PIASy (Chung et al. 1997). PIAS1 has been named as it has been shown to
inhibit STAT1, while PIAS3 interacts with STAT3 (Chung et al. 1997, Liu et al. 1998). Interest­
ingly, PIAS proteins neither act only as inhibitors nor show any specificity for STATs. PIAS can
regulate transcription either positively or negatively. The domain structure of PIAS include an
N-terminal SAP domain (SAF-A/B, Acinus, and PIAS), apoptotic chromatin-condensation
inducer in the nucleus, the PINIT (Pro-Ile-Asn-Ile-Thr) motif, a C3HC4-motif type RING finger-
like zinc-binding domain, an acidic amino acid domain (AD), SUMO-interacting motif, and
a variable C-terminal Ser/Thr-rich (S/T) domain (Aravind and Koonin 2000, Jackson 2001,
Jimenez-Lara, Heine, and Gronemeyer 2002, Duval et al. 2003). The SAP domain, the PINIT
motif, and the RING domain contribute to the nuclear localization of PIAS proteins (Duval et al.
2003). The SAP domain is important for sequence/structure-specific DNA binding (Aravind and
Koonin 2000). PIAS proteins could modulate JAK-STAT signaling by various means including
SUMOylation that attenuate transcription factor activity (Niu et al. 2018). The RING domain
might interact with other proteins, which is important for SUMO-E3-ligase function (Sachdev
et al. 2001). The function of AD and S/T domain remain less explored. SUMO ligase-independent
modulation of transcription factor activity of PIAS proteins exists (Kotaja et al. 2002). One such
mechanism is blocking the DNA-binding of a transcription factor. PIAS1 has been shown to
inhibit DNA-binding activities of STAT1 (Liu et al. 1998). Similarly, PIAS3 interaction with
STAT3 inhibited the DNA-binding activity of STAT3 (Chung et al. 1997). It has been hypothe­
sized that DNA-binding activity is inhibited by PIAS, obstructing the DNA-binding site of the
transcription factor within the PIAS: transcription factor complex. This could result in
a conformational change or dissociation of a transcription complex due to PIAS engagement of
transcription factor. Co-regulators, which can get recruited to a protein complex, are important
players for PIAS-mediated regulation of transcription. PIAS proteins interact with HDAC
molecules. Knockout mice studies have indicated functional redundancy between PIAS family
members. It is evident in PIASy-deficient mice that reveal minor defects in Wnt and IFN signaling
(Roth et al. 2004). Understanding the complex in vivo functions of PIAS proteins will depend on
multiple experimental approaches.

2.7.3 Effect of Suppressor of Cytokine Signaling


JAK-STAT pathway is also negatively regulated by suppressor of cytokine signaling proteins or
SOCS proteins. These proteins are not constitutively expressed. SOCS proteins bind to JAKs,
cytokine receptors, and signaling molecules. SOCS genes encode a family of eight proteins
(SOCS1-7 and cytokine-inducible SH2 protein (CIS)) (Endo et al. 1997, Naka et al. 1997, Starr
et al. 1997). These proteins contain an N-terminal of variable length, a central SH2 domain
that determines binding specificity, and a C-terminal termed as SOCS box (40 amino acid
module) that is implicated in proteasomal degradation (Endo et al. 1997, Naka et al. 1997,
Starr et al. 1997). CIS, SOCS1, 2, and 3 are well characterized, while the biological roles of
SOCS4–SOCS7 remain poorly understood (Hilton et al. 1998). CIS/SOCS family proteins act
as E3 ubiquitin ligase, mediating protein degradation of protein families associated with them.
A 24-amino acid kinase inhibitory region (KIR) is present in the N-terminal of SOCS1 and
SOCS3 (Nicholson et al. 1999). A series of studies on SOCS proteins have revealed that a wide
range of cytokines that utilize JAK-STAT signaling are inhibited by SOCS. The SH2 domain of
SOCS1 inhibits JAK2 by binding to Tyr1007 through insertion of its KIR into the activation
loop of the kinase domain (Yasukawa et al. 1999). This prevents substrates from getting into the
catalytic pocket by acting as pseudosubstrate. SOCS1 has also shown to abrogate tyrosine phosphor­
ylation of STAT1 via direct binding to phosphorylated type I IFN receptors (Fenner et al. 2006).
A model of SH2 domain: receptor interaction and KIR-mediated inhibition have been proposed.
However, the SH2 domain of SOCS3 does not possess a strong affinity toward activation loops of
JAKs (Yoshimura and Yasukawa 2012). The absence of SOCS1 expression due to gene methylation
results in the development of a range of primary cancers (Sutherland et al. 2004). SOCS3 first
20 JAK-STAT Signaling in Diseases

binds to the receptor, followed by inhibition of JAK activity via a KIR-mediated JAK inhibition
(Sasaki et al. 1999). SOCS3 also inhibits the signaling initiated by various stimuli that induce its
expression. This inhibition is predominantly achieved by SOCS3 engagement of the activated
receptors of these cytokines. SOCS3 has been demonstrated to specifically bind to phosphory­
lated tyrosines of cytokine/growth-factor receptors. A number of these receptor-docking sites also
act as binding sites for another signaling regulator, SHP2. The relative contribution of SOCS and
SHP2 to signal modulation and pathway crosstalk is still not completely understood. Intriguingly,
SOCS3 is not phosphorylated by JAKs, but relies on other kinases, such as Src kinases and
receptor-tyrosine kinases, for activation (Sommer et al. 2005). Absence of SOCS3 has been
implicated to enhance susceptibility to chronic inflammatory diseases. A lack of SOCS3 correlates
with the development and progression of malignancies (Riehle et al. 2008). The inhibitory
mechanisms of CIS and SOCS2 differ from that of SOCS1 and SOCS3 (Krebs and Hilton 2001).
Neither CIS nor SOCS2 acts on JAKs (Krebs and Hilton 2001). The biological role of CIS is not
completely understood. There might be functional redundancy of CIS. Role of SOCS2 has been
evaluated following the development of SOCS2-deficient mice that displays gigantism phenotype
rising from dysfunction of growth hormone axis (Metcalf et al. 2000). SOCS2 partially inhibits
recruitment of STAT5 by binding pTyr487 and pTyr595 on the growth hormone receptor
(Greenhalgh et al. 2005). A number of studies have linked SOCS2 to cancer, including colon and
prostate cancer.
SOCS are basically ubiquitin ligases that promote the degradation of their partners. The SOCS
box interacts with Elongin B and Elongin C, Cullin-5, and RING-box-2 to form an E3 ligase
complex termed as ECS (Elongin B/C-Cullin-SOCS box protein) (Bullock et al. 2006). The ECS
binds to a ubiquitin-activating enzyme (E1) and a ubiquitin-conjugating enzyme (E2), allowing
the ECS complex to recognize and bind target proteins for polyubiquitination and degradation
via the 26S proteasome (Kamura et al. 2004). Proteasomal degradation has emerged as a key
mechanism of signal attenuation. The nuclear protein, SLIM, has been characterized as
a ubiquitin E3 ligase that targets STATs for degradation via the proteasome (Tanaka, Soriano,
and Grusby 2005). The interaction of the SOCS box with the ECS complex provides a clear link
between SOCS proteins and proteasomal degradation that is central to attenuating the levels of
SOCS proteins within the cell (Zhang et al. 1999). It can be argued if SOCS proteins can act as
adaptors to bridge their binding partners to the proteasome and induce their degradation.
Various overexpression studies including SOCS proteins have supported a cross-talk between
different members of the SOCS family. Physiological evidence for the importance of the SOCS
box has been garnered from in vivo studies.

2.7.4 STAT Inhibitors as Therapeutic to Treat Inflammatory Disorders


JAK-STAT pathway is critical to the pathogenesis of inflammatory disorders mediated by the
immune system. Back in the 1990s, the potential of JAK inhibitors to treat autoimmune and
inflammatory disorders had been conceived (Banerjee et al. 2017). The role of JAK-STAT
signaling in various inflammatory disorders has been extensively mentioned in this book.
Extensive efforts have been generated to develop second-class JAK inhibitors that are more
selective. Dysregulated expression of STAT molecules has been observed in various cancers.
Hence, the discovery of STAT inhibitors has mainly targeted tumor cells. In the last section of
the chapter, we will discuss the development of STAT inhibitors to treat inflammatory disorders.
We will discuss the advent of small molecule inhibitors of STAT molecules for therapeutic
purposes.

2.7.4.1 STAT1 Inhibitor


Studies have demonstrated the development of very few STAT1 inhibitors. BRD0476, a novel
suppressor of pancreatic β-cell apoptosis, was developed. BRD0476 inhibited IFN-γ-induced
JAK-STAT Structure Function 21

JAK2 and STAT1 signaling to augment β-cell survival (Chou et al. 2015). BRD0476 inhibits
JAK-STAT signaling without suppressing the kinase activity of any JAK. An analog of BRD0476
was developed by adding quinolone moieties in place of naphthyl that led to more than 1000-fold
solubility as well inhibited IFN-γ/STAT1 signal transduction (Scully et al. 2013). Pravastatin is
another STAT1 inhibitor that attenuates the IFN-γ levels and prevents STAT1 phosphorylation
by inhibiting HMG-CoA reductase (Zhou, Gao, and Sun 2009). Pravastatin treatment on
Apolipoprotein E-deficient mice fed on a cholesterol-rich diet suppressed IFN-γ levels in both
serum and atherosclerotic lesions, associated with reduced STAT1 activation, attenuated IRF1
expression, and induced activity of SOCS1 in aorta tissue (Zhou, Gao, and Sun 2009). ISS-840,
a phosphopeptide mimetic of STAT1, disrupted STAT1 dimerization leading to moderately
potent inhibition of STAT1 signaling (Gunning et al. 2007). Fludarabine, a compound used to
treat hematologic malignancies, is another STAT1 activation inhibitor causing a specific depletion
of STAT1 protein, but not of other STAT molecules (Feng et al. 2017). Fludarabine exerts
apoptosis through increasing Bax and decreasing Bid, XIAP, and survivin expression (Meng
et al. 2007).

2.7.4.2 STAT3 Inhibitor


Among the STAT-specific inhibitors, STAT3 inhibitors are most widely studied. S3I-201, a novel
STAT3 inhibitor identified from NCI chemical library, displayed potent inhibition of STAT3 DNA-
binding activity with IC50 of 86 μM in cell-free assays (Siddiquee et al. 2007a). S3I-201 inhibited
STAT3 transcriptional activities, and induced growth inhibition and apoptosis of tumor cells
expressing constitutively active STAT3 (Grandis et al. 2000). In liver fibrosis, angiogenesis and
fibrogenesis were attenuated by S3I-201 (Wang et al. 2018). Derivatives of S3I-201, S3I-201.1066,
S3I-1757, and BP-1-102, displayed increased potencies of STAT3 inhibition (Zhang et al. 2010, 2012,
2013). S3I-201 administration resulted in appreciable decrease in IFN-γ, T-bet, IL-17A, RORγt,
Stat3, IL-21, and IL-22 levels, and increase in Foxp3 and Helios production CD4+ T cells in BTBR
mice, suggesting that S3I-201 could be used to treat autism (Ahmad et al. 2018). S3I-M2001 is an
oxazole-based peptidomimetic of the STAT3 SH2 domain-binding phosphotyrosine peptide that
selectively disrupts active STAT3:STAT3 dimers in various cancers where STAT3 expression is
aberrant (Siddiquee et al. 2007b). Stattic, a nonpeptidic small molecule identified after screening of
chemical libraries, has been displayed to selectively inhibit the function of the STAT3 SH2 domain at
10 μM, independent of the STAT3 activation state in vitro (Schust et al. 2006). STAT3 activation,
dimerization, and nuclear translocation are selectively inhibited by Stattic (Schust et al. 2006).
Fragments of STX-0119 (another STAT3 SH2 domain antagonist) and Stattic were chemically fused
to generate HJC0123, which suppressed STAT3 phosphorylation and transcriptional activity in
breast cancer cells and induced anti-tumor cell effects against breast and pancreatic cancer cells
in vitro at IC50 values of 0.1–1.25 μM (Ashizawa et al. 2011, Chen et al. 2013b). By inhibiting STAT3,
Stattic reduces the growth and increases the apoptosis of NPC and sensitizes NPC to cisplatin and IR
(Pan et al. 2013). STA-21 is another small molecule inhibitor of STAT3 with a potency of 20 μM.
STA-21 inhibits STAT3 dimerization and binding to DNA resulting in a transcriptional blockade of
STAT3-dependent genes (Song et al. 2005). STA-21 has been shown to be useful in several
inflammatory disease conditions (Park et al. 2014). STA-21 suppresses joint inflammation in
a mouse CIA model through regulation of the transcription factors, Foxp3 and RORγt, in CD8+
T cells and inhibits the development and progression of rheumatoid arthritis (Ahmad et al. 2017).
Studies suggest that LLL12, another small-molecule inhibitor of STAT3 signaling, might suppress
STAT3 activation by blocking its recruitment to the receptor, preventing phosphorylation by tyrosine
kinases, and by interfering with dimerization (Bid et al. 2012). LLL12 suppressed cell viability,
induced apoptosis, and repressed colony formation and migration in vitro in studies of glioblastoma,
osteosarcoma, and breast cancer cells (Lin et al. 2010, Onimoe et al. 2012). In a drug development
program involving virtual ligand screening, 2-D similarity screening, 3-D pharmacophore analysis,
and SAR-based medicinal chemistry, C188-9, a potent small-molecule that targets the Src-homology
22 JAK-STAT Signaling in Diseases

SH2 domain of STAT3 was identified (Jung et al. 2017). C188-9 inhibited growth and survival of
several cancer cell lines in vitro, including breast cancer, acute myeloid leukemia, head and neck
squamous cell carcinoma, and non–small cell lung cancer (Jung et al. 2017). C188-9 impedes nuclear
translocation of STAT3, prevents STAT3 binding to its target gene promoters, and inhibits STAT3­
mediated regulation of gene expression (Bharadwaj et al. 2016). Another small molecule, non­
phosphorylated STAT3 inhibitor, SH-4-54 strongly binds to STAT3 protein (KD = 300 nM) (166).
SH-4-54 potently kills glioblastoma brain cancer stem cells and suppresses STAT3 phosphorylation
and its downstream transcriptional targets at low concentrations in the nM level (Haftchenary et al.
2013). Moreover, in vivo, the inhibitor exhibited blood–brain barrier permeability, potently controlled
glioma tumor growth, and inhibited pSTAT3 in vivo (Haftchenary et al. 2013). HJC0152, an
O-alkylamino-tethered derivative of niclosamide is a potent STAT3 inhibitor with remarkably
improved aqueous solubility (Chen et al. 2013a). HJC0152 exerted a significant anticancer effect on
HNSCC tumor growth and invasion (Wang et al. 2017). HJC0152 also inhibits STAT3 activation in
GC cells, and retards their growth in vitro and in vivo (Jiang et al. 2018). BBI608 is a small molecule
STAT3 inhibitor known to suppress cancer relapse, progression, and metastasis (MacDonagh et al.
2018). BBI608 can inhibit stemness gene expression, deplete CSCs, and overcome cisplatin resistance
in NSCLC (MacDonagh et al. 2018). BBI608 was initially introduced and investigated as a stemness
inhibitor in the context of tumor relapse in a xenograft model of pancreatic cancer. BBI608 was later
investigated in the context of prostate cancer progression. BBI608 inhibited cell proliferation, colony
formation, and migration, while increasing the sensitivity of prostate cancer cells to the cytotoxic
effects of docetaxel (Zhang et al. 2016). HO-3867 is an analog of curcumin that selectively suppresses
STAT3 phosphorylation, transcription, and DNA binding without affecting the expression of other
active STATs (Rath et al. 2014). This analog has been shown to induce apoptosis in BRCA-mutated
ovarian cancer cells with minimal toxicity to normal cells (Tierney et al. 2012). When combined with
cisplatin, HO-3867 has demonstrated synergistic inhibition of chemotherapy-resistant ovarian xeno­
graft tumors (174). HO-3867/cisplatin combination treatment significantly inhibited cisplatin­
resistant cell proliferation in a concentration-dependent manner (Selvendiran et al. 2011). HO-3867
induces cell apoptosis by reactive oxygen species-dependent endoplasmic reticulum stress in human
pancreatic cancer cells (Hu et al. 2017). APTSTAT3-9R is an example of specific STAT3-binding
peptide with addition of a cell-penetrating motif (Kim et al. 2014). STAT3-specific aptide APTSTAT3
has been developed for therapeutic efficacy in cancer by inhibiting STAT3 signaling (Kim et al. 2014).
The treatment of APTSTAT3-9R in various types of cancer cells blocks STAT3 phosphorylation and
reduces expression of STAT targets (Kim et al. 2014). Peptide-based STAT3 inhibitors have also been
reported. These inhibitory phospho-peptides (PYLKTK, YLPQTV), based on a shortened gp130
sequence, bind to the SH2 domain of STAT3 and block dimerization (Zhang et al. 2010). In an in vivo
setting, however, the relatively low potency and limited cell permeability of such phospho-peptides
have hampered their further use. Peptide libraries have been screened for interaction with the DNA-
binding domain (DBD) of STAT3, and the identified peptides have been shown to inhibit STAT3
DNA-binding activity (Kim et al. 2014). ISS-610 inhibited malignant cell growth and induced
apoptosis in vitro. Peptidomimetic ISS-610 also exhibited good selectivity against STAT3 and its
functions (Turkson et al. 2004). ISS-610 with 4-cyanobenzoate substitution inhibits constitutive
STAT3 activity in Src-transformed mouse fibroblasts and human breast and lung cancer cells
(Turkson et al. 2004). CJ-1383 is another class of small molecule inhibitors that targets STAT3-SH2
domain. CJ-1383 binds to STAT3 with a Ki value of 0.95 μM and inhibits cell growth with IC50 values
of 3−11 μM in two breast cancer cell lines having increased phospho-STAT3 expression (Chen et al.
2010). In a time-and dose-dependent fashion, CJ-1383 is effective in inhibition of STAT3 activity and
induction of apoptosis in the MDA-MB-468 cancer cell line (Chen et al. 2010). PM-73G is one of the
first phosphopeptides targeted to an SH2 domain that inhibits its target in vivo following systemic
administration (Auzenne et al. 2012). PM-73G possesses the non-hydrolyzable difluoromethylpho­
sphonate group, and the esterase-labile POM groups block the negative charges thus allowing passive
diffusion across cell membranes (Auzenne et al. 2012). Intratumoral and intraperitoneal administra­
tion of PM-73G to mice-bearing MDA-MB-468 tumor xenografts demonstrate significant inhibition
of tumor growth in vivo, despite having no discernible in vitro effect (Auzenne et al. 2012). Zhao et al.
JAK-STAT Structure Function 23

designed a non-peptide small molecule STAT3 inhibitor, LY5, using in silico site-directed fragment-
based drug design (FBDD) (Zhao et al. 2016). The inhibitory effect on STAT3 phosphorylation, cell
viability, migration, and colony-forming ability by LY5 were examined in human liver and colon
cancer cells (Zhao et al. 2016). LY5-inhibited constitutive IL-6-induced STAT3 phosphorylation,
STAT3 nuclear translocation, decreased STAT3 downstream targeted gene expression, and induced
apoptosis in liver and colon cancer cells (Zhao et al. 2016). LY5 had little effect on STAT1
phosphorylation mediated by IFN-γ (Zhao et al. 2016).

2.7.4.3 STAT4 Inhibitor


Most STAT4 inhibitors are natural products. STAT4 protein levels decrease in response to
berbamine treatment (Ren et al. 2008). Studies also showed that the oral administration of water
extract of cinnamon bark decreased IFN-y expression and inhibited STAT4 activation in
activated murine T cells (Lee et al. 2011). Tofacitinib has been shown to inhibit STAT4 activation
in cultured anti-CD3-stimulated T cells (Migita et al. 2011).

2.7.4.4 STAT5 Inhibitor


SH-4-54 that potently inhibited STAT3 also inhibits STAT5 with KD value of 464 nM (Haftchenary
et al. 2013). The non-peptide small molecule pimozide, a neuroleptic drug, inhibits the constitutive
STAT5 Tyr694 phosphorylation (5–10 µM for 3 h) and transcription activity (5 µM for 18 h) in Bcr-
Abl+ K562 and KU812 cultures (Nelson et al. 2011). Pimozide inhibits the activating tyrosine
phosphorylation of STAT5 (Nelson et al. 2011). However, several lines of evidence suggest that
pimozide does not function as a kinase inhibitor. Pimozide neither attenuates the autophosphoryla­
tion of BCR/ABL, nor does it decrease the tyrosine phosphorylation of other cellular proteins
mediated by BCR/ABL (Nelson et al. 2011). In addition, pimozide does not decrease the activation
of a distinct pathway downstream of BCR/ABL, MAP kinase (Nelson et al. 2011). When Berg et al.
screened large small-molecule libraries in search of compounds that can modulate SH2 domain of
STAT5, chromone nicotinyl hydrazone was discovered (Muller et al. 2008). The discovery of
chromone aldehydes as inhibitors of STAT family proteins indicates toward a model by which,
under the assay conditions, acyl hydrazones are cleaved to the respective aldehydes representing the
active species. It shows reduced potency toward the SH2 domains of STAT1, STAT3, or LCK but
suppresses SH2 domain of STAT5 at IC50 of 47 μM (Muller et al. 2008). STAT5 is activated by FLT3­
ITD, which is a constitutively active TK driving the pathogenesis of AML (Wingelhofer et al. 2018).
Wingelhofer et al. developed a novel, potent STAT5 SH2 domain inhibitor, AC-4–130, which can
efficiently block pathological levels of STAT5 activity in AML (Wingelhofer et al. 2018). AC-4–130
directly binds to STAT5, thereby disrupting STAT5 activation, dimerization, nuclear translocation,
and STAT5-dependent gene transcription. AC-4-130 shows high selectivity for STAT5 over STAT1
and STAT3 (Cumaraswamy et al. 2014, Wingelhofer et al. 2018).

2.7.4.5 STAT6 Inhibitor


There are small molecule inhibitors of STAT6. AS1517499 is a novel selective STAT6 inhibitor
with an IC50 value of 21 nM (Nagashima et al. 2007). AS1517499 is a potent STAT6 inhibitor
synthesized based on the structure of a reported STAT6 inhibitor, TMC-264, discovered from
the fungus Phoma (Sakurai et al. 2003). AS1517499 also inhibited the IL-4–induced Th2 cell
differentiation of mouse spleen T cells with an IC50 value of 2.3 nM without influencing the
IL-12–induced Th1 cell differentiation (Sakurai et al. 2003). In vivo treatment with AS1517499
ameliorates the antigen-induced bronchial smooth muscle (BSM) hyperresponsiveness by
inhibiting the RhoA up-regulation in BSMs and, at least in part, by reducing the IL-13
production in the airways in mice (Chiba et al. 2009). TMC-264 is known to inhibit both
tyrosine phosphorylation of STAT6, with an IC50 value of 1.6 mM, and the complex
24 JAK-STAT Signaling in Diseases

formation of phosphorylated STAT6 with its recognition DNA sequence (Sakurai et al. 2003).
A novel inhibitor of STAT6, PM-242H, inhibited initiation of allergic disease induced by
airway fungal challenge, reversed established allergic airway disease in mice, and blocked
salmeterol-dependent enhanced allergic airway disease (Knight et al. 2015). In vitro structure–
activity relationship studies led to the development of the lead compound, PM-43I (Knight
et al. 2018). Conducting initial dose range, toxicity, and pharmacokinetic experiments with
PM-43I potently inhibits both STAT5- and STAT6-dependent allergic airway diseases in mice
(Knight et al. 2018). When cellular activity screen was performed for evaluating the binding
capacity to SH2 domains of STAT molecules, PM-86I showed the highest specificity for STAT6
and no cross-reactivity to any of the additional targets at the highest dose tested (Knight et al.
2018). PM-86I can specifically inhibit STAT6-dependent adaptive Th2 immune responses when
administered systemically (Knight et al. 2018). Other notable inhibitors of STAT6 signaling
include AS1810722, a fused bicyclic pyrimidine derivative (IC50value of 2.4 nM) (Nagashima
et al. 2009). AS1810722 is potent, orally active STAT6 inhibitors that could be developed for the
treatment of STAT6-dependent allergic diseases, such as asthma. AS1810722 showed potent
STAT6 inhibition and a good CYP3A4 inhibition profile (Nagashima et al. 2009).

2.8 Concluding Remarks


JAK-STAT is one of the well-characterized signaling pathways involved in maintaining home­
ostasis. Aberrant regulation of this pathway affects the downstream signaling molecules leading to
deleterious consequences including onset and pathophysiology of auto-immune disorders. This
chapter has elaborated on the structure of JAK and STAT domains that can be targeted for
therapeutic benefits. Subsequent chapters of the book have discussed the implication of JAK­
STAT signaling in various inflammatory disorders.

REFERENCES
Ahmad, S. F., M. A. Ansari, A. Nadeem, S. A. Bakheet, M. A. Alshammari, M. R. Khan,
A. M. S. Alsaad, and S. M. Attia. 2018. “S3I-201, a selective Stat3 inhibitor, restores neuroimmune
function through upregulation of Treg signaling in autistic BTBR T(+) Itpr3(tf)/J mice.” Cell Signal
52:127–36. doi: 10.1016/j.cellsig.2018.09.006.
Ahmad, S. F., M. A. Ansari, A. Nadeem, K. M. A. Zoheir, S. A. Bakheet, A. M. S. Alsaad,
O. A. Al-Shabanah, and S. M. Attia. 2017. “STA-21, a STAT-3 inhibitor, attenuates the
development and progression of inflammation in collagen antibody-induced arthritis.” Immunobiology
222 (2):206–17. doi: 10.1016/j.imbio.2016.10.001.
Akira, S., Y. Nishio, M. Inoue, X. J. Wang, S. Wei, T. Matsusaka, K. Yoshida, T. Sudo, M. Naruto, and
T. Kishimoto. 1994. “Molecular cloning of APRF, a novel IFN-stimulated gene factor 3 p91-related
transcription factor involved in the gp130-mediated signaling pathway.” Cell 77 (1):63–71.
Alonso, A., J. Sasin, N. Bottini, I. Friedberg, I. Friedberg, A. Osterman, A. Godzik, T. Hunter, J. Dixon,
and T. Mustelin. 2004. “Protein tyrosine phosphatases in the human genome.” Cell 117 (6):699–711.
doi: 10.1016/j.cell.2004.05.018.
Aravind, L., and E. V. Koonin. 2000. “SAP – a putative DNA-binding motif involved in chromosomal
organization.” Trends Biochem Sci 25 (3):112–14.
Ardito, F., M. Giuliani, D. Perrone, G. Troiano, and L. Lo Muzio. 2017. “The crucial role of protein
phosphorylation in cell signaling and its use as targeted therapy (review).” Int J Mol Med 40
(2):271–80. doi: 10.3892/ijmm.2017.3036.
Argetsinger, L. S., J. L. Kouadio, H. Steen, A. Stensballe, O. N. Jensen, and C. Carter-Su. 2004.
“Autophosphorylation of JAK2 on tyrosines 221 and 570 regulates its activity.” Mol Cell Biol 24
(11):4955–67. doi: 10.1128/MCB.24.11.4955-4967.2004.
Ashizawa, T., H. Miyata, H. Ishii, C. Oshita, K. Matsuno, Y. Masuda, T. Furuya, T. Okawara,
M. Otsuka, N. Ogo, A. Asai, and Y. Akiyama. 2011. “Antitumor activity of a novel small molecule
JAK-STAT Structure Function 25

STAT3 inhibitor against a human lymphoma cell line with high STAT3 activation.” Int J Oncol 38
(5):1245–52. doi: 10.3892/ijo.2011.957.
Auzenne, E. J., J. Klostergaard, P. K. Mandal, W. S. Liao, Z. Lu, F. Gao, R. C. Bast, Jr., F. M. Robertson,
and J. S. McMurray. 2012. “A phosphopeptide mimetic prodrug targeting the SH2 domain of Stat3
inhibits tumor growth and angiogenesis.” J Exp Ther Oncol 10 (2):155–62.
Babon, J. J., I. S. Lucet, J. M. Murphy, N. A. Nicola, and L. N. Varghese. 2014. “The molecular regulation
of Janus kinase (JAK) activation.” Biochem J 462 (1):1–13. doi: 10.1042/BJ20140712.
Baker, S. J., S. G. Rane, and E. P. Reddy. 2007. “Hematopoietic cytokine receptor signaling.” Oncogene 26
(47):6724–37. doi: 10.1038/sj.onc.1210757.
Banerjee, S., A. Biehl, M. Gadina, S. Hasni, and D. M. Schwartz. 2017. “JAK–STAT signaling as a target
for inflammatory and autoimmune diseases: current and future prospects.” Drugs 77 (5):521–46.
doi: 10.1007/s40265-017-0701-9.
Becker, S., B. Groner, and C. W. Muller. 1998. “Three-dimensional structure of the Stat3beta homodimer
bound to DNA.” Nature 394 (6689):145–51. doi: 10.1038/28101.
Bharadwaj, U., T. K. Eckols, X. Xu, M. M. Kasembeli, Y. Chen, M. Adachi, Y. Song, Q. Mo, S. Y. Lai,
and D. J. Tweardy. 2016. “Small-molecule inhibition of STAT3 in radioresistant head and neck
squamous cell carcinoma.” Oncotarget 7 (18):26307–30. doi: 10.18632/oncotarget.8368.
Bid, H. K., D. Oswald, C. Li, C. A. London, J. Lin, and P. J. Houghton. 2012. “Anti-angiogenic activity of
a small molecule STAT3 inhibitor LLL12.” PLoS One 7 (4):e35513. doi: 10.1371/journal.pone.0035513.
Boggon, T. J., Y. Li, P. W. Manley, and M. J. Eck. 2005. “Crystal structure of the Jak3 kinase
domain in complex with a staurosporine analog.” Blood 106 (3):996–1002. doi: 10.1182/blood­
2005-02-0707.
Bohmer, F. D., and K. Friedrich. 2014. “Protein tyrosine phosphatases as wardens of STAT signaling.”
JAKSTAT 3 (1):e28087. doi: 10.4161/jkst.28087.
Bourdeau, A., N. Dube, and M. L. Tremblay. 2005. “Cytoplasmic protein tyrosine phosphatases,
regulation and function: the roles of PTP1B and TC-PTP.” Curr Opin Cell Biol 17 (2):203–09. doi:
10.1016/j.ceb.2005.02.001.
Bousoik, E., and H. Montazeri Aliabadi. 2018. ““Do we know Jack” about JAK? A closer look at
JAK–STAT signaling pathway.” Front Oncol 8:287. doi: 10.3389/fonc.2018.00287.
Bradshaw, J. M., V. Mitaxov, and G. Waksman. 1999. “Investigation of phosphotyrosine recognition by
the SH2 domain of the Src kinase.” J Mol Biol 293 (4):971–85. doi: 10.1006/jmbi.1999.3190.
Bradshaw, J. M., and G. Waksman. 2002. “Molecular recognition by SH2 domains.” Adv Protein Chem
61:161–210.
Bullock, A. N., J. E. Debreczeni, A. M. Edwards, M. Sundstrom, and S. Knapp. 2006. “Crystal structure
of the SOCS2-elongin C-elongin B complex defines a prototypical SOCS box ubiquitin ligase.” Proc
Natl Acad Sci U S A 103 (20):7637–42. doi: 10.1073/pnas.0601638103.
Chen, H., Z. Yang, C. Ding, L. Chu, Y. Zhang, K. Terry, H. Liu, Q. Shen, and J. Zhou. 2013a. “Discovery
of O-alkylamino tethered niclosamide derivatives as potent and orally bioavailable anticancer
agents.” ACS Med Chem Lett 4 (2):180–85. doi: 10.1021/ml3003082.
Chen, H., Z. Yang, C. Ding, L. Chu, Y. Zhang, K. Terry, H. Liu, Q. Shen, and J. Zhou. 2013b. “Fragment­
based drug design and identification of HJC0123, a novel orally bioavailable STAT3 inhibitor for
cancer therapy.” Eur J Med Chem 62:498–507. doi: 10.1016/j.ejmech.2013.01.023.
Chen, J., L. Bai, D. Bernard, Z. Nikolovska-Coleska, C. Gomez, J. Zhang, H. Yi, and S. Wang. 2010.
“Structure-based design of conformationally constrained, cell-permeable STAT3 inhibitors.” ACS
Med Chem Lett 1 (2):85–89. doi: 10.1021/ml100010j.
Chiba, Y., M. Todoroki, Y. Nishida, M. Tanabe, and M. Misawa. 2009. “A novel STAT6 inhibitor
AS1517499 ameliorates antigen-induced bronchial hypercontractility in mice.” Am J Respir Cell
Mol Biol 41 (5):516–24. doi: 10.1165/rcmb.2008-0163OC.
Chou, D. H., A. Vetere, A. Choudhary, S. S. Scully, M. Schenone, A. Tang, R. Gomez, S. M. Burns,
M. Lundh, T. Vital, E. Comer, P. W. Faloon, V. Dancik, C. Ciarlo, J. Paulk, M. Dai, C. Reddy,
H. Sun, M. Young, N. Donato, J. Jaffe, P. A. Clemons, M. Palmer, S. A. Carr, S. L. Schreiber, and
B. K. Wagner. 2015. “Kinase-independent small-molecule inhibition of JAK–STAT signaling.”
J Am Chem Soc 137 (24):7929–34. doi: 10.1021/jacs.5b04284.
Chung, C. D., J. Liao, B. Liu, X. Rao, P. Jay, P. Berta, and K. Shuai. 1997. “Specific inhibition of Stat3
signal transduction by PIAS3.” Science 278 (5344):1803–05. doi: 10.1126/science.278.5344.1803.
26 JAK-STAT Signaling in Diseases

Cornejo, M. G., T. J. Boggon, and T. Mercher. 2009. “JAK3: a two-faced player in hematological
disorders.” Int J Biochem Cell Biol 41 (12):2376–79. doi: 10.1016/j.biocel.2009.09.004.
Cumaraswamy, A. A., A. M. Lewis, M. Geletu, A. Todic, D. B. Diaz, X. R. Cheng, C. E. Brown,
R. C. Laister, D. Muench, K. Kerman, H. L. Grimes, M. D. Minden, and P. T. Gunning. 2014.
“Nanomolar-potency small molecule inhibitor of STAT5 protein.” ACS Med Chem Lett 5
(11):1202–06. doi: 10.1021/ml500165r.
Decker, T., and P. Kovarik. 2000. “Serine phosphorylation of STATs.” Oncogene 19 (21):2628–37. doi:
10.1038/sj.onc.1203481.
Delgoffe, G. M., and D. A. Vignali. 2013. “STAT heterodimers in immunity: a mixed message or a unique
signal?” JAKSTAT 2 (1):e23060. doi: 10.4161/jkst.23060.
Duval, D., G. Duval, C. Kedinger, O. Poch, and H. Boeuf. 2003. “The ‘PINIT’ motif, of a newly identified
conserved domain of the PIAS protein family, is essential for nuclear retention of PIAS3L.” FEBS
Lett 554 (1–2):111–18. doi: 10.1016/s0014-5793(03)01116-5.
Endo, T. A., M. Masuhara, M. Yokouchi, R. Suzuki, H. Sakamoto, K. Mitsui, A. Matsumoto,
S. Tanimura, M. Ohtsubo, H. Misawa, T. Miyazaki, N. Leonor, T. Taniguchi, T. Fujita,
Y. Kanakura, S. Komiya, and A. Yoshimura. 1997. “A new protein containing an SH2 domain
that inhibits JAK kinases.” Nature 387 (6636):921–24. doi: 10.1038/43213.
Feng, Z., W. Zheng, Q. Tang, L. Cheng, H. Li, W. Ni, and X. Pan. 2017. “Fludarabine inhibits
STAT1-mediated up-regulation of caspase-3 expression in dexamethasone-induced osteoblasts
apoptosis and slows the progression of steroid-induced avascular necrosis of the femoral head in
rats.” Apoptosis 22 (8):1001–12. doi: 10.1007/s10495-017-1383-1.
Fenner, J. E., R. Starr, A. L. Cornish, J. G. Zhang, D. Metcalf, R. D. Schreiber, K. Sheehan, D. J. Hilton,
W. S. Alexander, and P. J. Hertzog. 2006. “Suppressor of cytokine signaling 1 regulates the immune
response to infection by a unique inhibition of type I interferon activity.” Nat Immunol 7 (1):33–39.
doi: 10.1038/ni1287.
Ferrao, R., and P. J. Lupardus. 2017. “The janus kinase (JAK) FERM and SH2 domains: bringing
specificity to JAK-receptor interactions.” Front Endocrinol (Lausanne) 8:71. doi: 10.3389/
fendo.2017.00071.
Ferrao, R., H. J. Wallweber, H. Ho, et al. 2016. “The structural basis for class II cytokine receptor
recognition by JAK1.” Structure 24 (6):897–905. doi: 10.1016/j.str.2016.03.023.
Filippakopoulos, P., S. Muller, and S. Knap. 2009. “SH2 domains: modulators of nonreceptor tyrosine
kinase activity.” Curr Opin Struct Biol 19 (6):643–49. doi: 10.1016/j.sbi.2009.10.001.
Fischer, E. H., H. Charbonneau, and N. K. Tonks. 1991. “Protein tyrosine phosphatases: a diverse family
of intracellular and transmembrane enzymes.” Science 253 (5018):401–06. doi: 10.1126/
science.1650499.
Fornerod, M., M. Ohno, M. Yoshida, and I. W. Mattaj. 1997. “CRM1 is an export receptor for
leucine-rich nuclear export signals.” Cell 90 (6):1051–60.
Frame, M. C., H. Patel, B. Serrels, D. Lietha, and M. J. Eck. 2010. “The FERM domain: organizing the
structure and function of FAK.” Nat Rev Mol Cell Biol 11 (11):802–14. doi: 10.1038/nrm2996.
Frangioni, J. V., P. H. Beahm, V. Shifrin, C. A. Jost, and B. G. Neel. 1992. “The nontransmembrane
tyrosine phosphatase PTP-1B localizes to the endoplasmic reticulum via its 35 amino acid
C-terminal sequence.” Cell 68 (3):545–60. doi: 10.1016/0092-8674(92)90190-n.
Goldfarb, D. S., A. H. Corbett, D. A. Mason, M. T. Harreman, and S. A. Adam. 2004. “Importin alpha:
a multipurpose nuclear-transport receptor.” Trends Cell Biol 14 (9):505–14. doi: 10.1016/j.
tcb.2004.07.016.
Gorlich, D., and I. W. Mattaj. 1996. “Nucleocytoplasmic transport.” Science 271 (5255):1513–18. doi:
10.1126/science.271.5255.1513.
Grandis, J. R., S. D. Drenning, Q. Zeng, S. C. Watkins, M. F. Melhem, S. Endo, D. E. Johnson, L. Huang,
Y. He, and J. D. Kim. 2000. “Constitutive activation of Stat3 signaling abrogates apoptosis in
squamous cell carcinogenesis in vivo.” Proc Natl Acad Sci U S A 97 (8):4227–32. doi: 10.1073/
pnas.97.8.4227.
Greenhalgh, C. J., E. Rico-Bautista, M. Lorentzon, A. L. Thaus, P. O. Morgan, T. A. Willson,
P. Zervoudakis, D. Metcalf, I. Street, N. A. Nicola, A. D. Nash, L. J. Fabri, G. Norstedt,
C. Ohlsson, A. Flores-Morales, W. S. Alexander, and D. J. Hilton. 2005. “SOCS2 negatively
JAK-STAT Structure Function 27

regulates growth hormone action in vitro and in vivo.” J Clin Invest 115 (2):397–406. doi: 10.1172/
JCI22710.
Gunning, P. T., W. P. Katt, M. Glenn, K. Siddiquee, J. S. Kim, R. Jove, S. M. Sebti, J. Turkson, and
A. D. Hamilton. 2007. “Isoform selective inhibition of STAT1 or STAT3 homo-dimerization via
peptidomimetic probes: structural recognition of STAT SH2 domains.” Bioorg Med Chem Lett 17
(7):1875–78. doi: 10.1016/j.bmcl.2007.01.077.
Haftchenary, S., H. A. Luchman, A. O. Jouk, A. J. Veloso, B. D. Page, X. R. Cheng, S. S. Dawson,
N. Grinshtein, V. M. Shahani, K. Kerman, D. R. Kaplan, C. Griffin, A. M. Aman, R. Al-Awar,
S. Weiss, and P. T. Gunning. 2013. “Potent targeting of the STAT3 protein in brain cancer stem cells:
a promising route for treating glioblastoma.” ACS Med Chem Lett 4 (11):1102–07. doi: 10.1021/
ml4003138.
Hammaren, H. M., D. Ungureanu, J. Grisouard, R. C. Skoda, S. R. Hubbard, and O. Silvennoinen. 2015.
“ATP binding to the pseudokinase domain of JAK2 is critical for pathogenic activation.” Proc Natl
Acad Sci U S A 112 (15):4642–47. doi: 10.1073/pnas.1423201112.
Hendriks, W. J., A. Elson, S. Harroch, and A. W. Stoker. 2008. “Protein tyrosine phosphatases: functional
inferences from mouse models and human diseases.” Febs J 275 (5):816–30. doi: 10.1111/j.1742­
4658.2008.06249.x.
Hilton, D. J., R. T. Richardson, W. S. Alexander, E. M. Viney, T. A. Willson, N. S. Sprigg, R. Starr,
S. E. Nicholson, D. Metcalf, and N. A. Nicola. 1998. “Twenty proteins containing a C-terminal
SOCS box form five structural classes.” Proc Natl Acad Sci U S A 95 (1):114–19. doi: 10.1073/
pnas.95.1.114.
Hoey, T., and U. Schindler. 1998. “STAT structure and function in signaling.” Curr Opin Genet Dev 8
(5):582–87.
Hof, P., S. Pluskey, S. Dhe-Paganon, M. J. Eck, and S. E. Shoelson. 1998. “Crystal structure of the
tyrosine phosphatase SHP-2.” Cell 92 (4):441–50. doi: 10.1016/s0092-8674(00)80938-1.
Horvath, C. M., Z. Wen, and J. E. Darnell, Jr. 1995. “A STAT protein domain that determines DNA
sequence recognition suggests a novel DNA-binding domain.” Genes Dev 9 (8):984–94. doi:
10.1101/gad.9.8.984.
Hu, Y., C. Zhao, H. Zheng, K. Lu, D. Shi, Z. Liu, X. Dai, Y. Zhang, X. Zhang, W. Hu, and G. Liang.
2017. “A novel STAT3 inhibitor HO-3867 induces cell apoptosis by reactive oxygen
species-dependent endoplasmic reticulum stress in human pancreatic cancer cells.” Anticancer
Drugs 28 (4):392–400. doi: 10.1097/CAD.0000000000000470.
Ibarra-Sanchez, M. J., J. Wagner, M. T. Ong, C. Lampron, and M. L. Tremblay. 2001. “Murine embryonic
fibroblasts lacking TC-PTP display delayed G1 phase through defective NF-kappaB activation.”
Oncogene 20 (34):4728–39. doi: 10.1038/sj.onc.1204648.
Isaacs, A., and J. Lindenmann. 1957. “Virus interference. I. The interferon.” Proc R Soc Lond B Biol Sci
147 (927):258–67.
Ishida-Takahashi, R., F. Rosario, Y. Gong, K. Kopp, Z. Stancheva, X. Chen, E. P. Feener, and
M. G. Myers, Jr. 2006. “Phosphorylation of Jak2 on Ser(523) inhibits Jak2-dependent leptin
receptor signaling.” Mol Cell Biol 26 (11):4063–73. doi: 10.1128/MCB.01589-05.
Iversen, L. F., K. B. Moller, A. K. Pedersen, G. H. Peters, A. S. Petersen, H. S. Andersen, S. Branner,
S. B. Mortensen, and N. P. Moller. 2002. “Structure determination of T cell protein-tyrosine
phosphatase.” J Biol Chem 277 (22):19982–90. doi: 10.1074/jbc.M200567200.
Jackson, P. K. 2001. “A new RING for SUMO: wrestling transcriptional responses into nuclear bodies
with PIAS family E3 SUMO ligases.” Genes Dev 15 (23):3053–58. doi: 10.1101/gad.955501.
Jacobson, N. G., S. J. Szabo, R. M. Weber-Nordt, Z. Zhong, R. D. Schreiber, J. E. Darnell, Jr., and
K. M. Murphy. 1995. “Interleukin 12 signaling in T helper type 1 (Th1) cells involves tyrosine
phosphorylation of signal transducer and activator of transcription (Stat)3 and Stat4.” J Exp Med
181 (5):1755–62. doi: 10.1084/jem.181.5.1755.
Jans, D. A., C. Y. Xiao, and M. H. Lam. 2000. “Nuclear targeting signal recognition: a key control point
in nuclear transport?” Bioessays 22 (6):532–44. doi: 10.1002/(SICI)1521-1878(200006)22:6<532::
AID-BIES6>3.0.CO;2-O.
Jatiani, S. S., S. J. Baker, L. R. Silverman, and E. P. Reddy. 2010. “Jak/STAT pathways in cytokine
signaling and myeloproliferative disorders: approaches for targeted therapies.” Genes Cancer 1
(10):979–93. doi: 10.1177/1947601910397187.
28 JAK-STAT Signaling in Diseases

Jiang, X., M. Wu, Z. Xu, H. Wang, H. Wang, X. Yu, Z. Li, and L. Teng. 2018. “HJC0152, a novel STAT3
inhibitor with promising anti-tumor effect in gastric cancer.” Cancer Manag Res 10:6857–67. doi:
10.2147/CMAR.S188364.
Jimenez-Lara, A. M., M. J. Heine, and H. Gronemeyer. 2002. “PIAS3 (protein inhibitor of activated
STAT-3) modulates the transcriptional activation mediated by the nuclear receptor coactivator
TIF2.” FEBS Lett 526 (1–3):142–46. doi: 10.1016/s0014-5793(02)03154-x.
Jin, Y. J., C. L. Yu, and S. J. Burakoff. 1999. “Human 70-kDa SHP-1L differs from 68-kDa SHP-1 in its
C-terminal structure and catalytic activity.” J Biol Chem 274 (40):28301–7. doi: 10.1074/
jbc.274.40.28301.
Jung, K. H., W. Yoo, H. L. Stevenson, D. Deshpande, H. Shen, M. Gagea, S. Y. Yoo, J. Wang,
T. K. Eckols, U. Bharadwaj, D. J. Tweardy, and L. Beretta. 2017. “Multifunctional effects of a
small-molecule STAT3 inhibitor on NASH and hepatocellular carcinoma in mice.” Clin Cancer Res
23 (18):5537–46. doi: 10.1158/1078-0432.CCR-16-2253.
Kamatkar, S., V. Radha, S. Nambirajan, R. S. Reddy, and G. Swarup. 1996. “Two splice variants of
a tyrosine phosphatase differ in substrate specificity, DNA binding, and subcellular location.” J Biol
Chem 271 (43):26755–61. doi: 10.1074/jbc.271.43.26755.
Kamura, T., K. Maenaka, S. Kotoshiba, M. Matsumoto, D. Kohda, R. C. Conaway, J. W. Conaway, and
K. I. Nakayama. 2004. “VHL-box and SOCS-box domains determine binding specificity for
Cul2-Rbx1 and Cul5-Rbx2 modules of ubiquitin ligases.” Genes Dev 18 (24):3055–65. doi:
10.1101/gad.1252404.
Kaneko, T., R. Joshi, S. M. Feller, and S. S. Li. 2012. “Phosphotyrosine recognition domains: the typical,

the atypical and the versatile.” Cell Commun Signal 10 (1):32. doi: 10.1186/1478-811X-10-32.

Khwaja, A. 2006. “The role of Janus kinases in haemopoiesis and haematological malignancy.” Br

J Haematol 134 (4):366–84. doi: 10.1111/j.1365-2141.2006.06206.x.


Kile, B. T., B. A. Schulman, W. S. Alexander, N. A. Nicola, H. M. Martin, and D. J. Hilton. 2002. “The
SOCS box: a tale of destruction and degradation.” Trends Biochem Sci 27 (5):235–41.
Kim, D., I. H. Lee, S. Kim, M. Choi, H. Kim, S. Ahn, P. E. Saw, H. Jeon, Y. Lee, and S. Jon. 2014.
“A specific STAT3-binding peptide exerts antiproliferative effects and antitumor activity by
inhibiting STAT3 phosphorylation and signaling.” Cancer Res 74 (8):2144–51. doi: 10.1158/0008­
5472.CAN-13-2187.
Kisseleva, T., S. Bhattacharya, J. Braunstein, and C. W. Schindler. 2002. “Signaling through the JAK–
STAT pathway, recent advances and future challenges.” Gene 285 (1–2):1–24.
Knight, J. M., G. Mak, J. Shaw, P. Porter, C. McDermott, L. Roberts, R. You, X. Yuan,
V. O. Millien, Y. Qian, L. Z. Song, V. Frazier, C. Kim, J. J. Kim, R. A. Bond, J. D. Milner,
Y. Zhang, P. K. Mandal, A. Luong, F. Kheradmand, J. S. McMurray, and D. B. Corry. 2015.
“Long-acting beta agonists enhance allergic airway disease.” PLoS One 10 (11):e0142212. doi:
10.1371/journal.pone.0142212.
Knight, J. M., P. Mandal, P. Morlacchi, G. Mak, E. Li, M. Madison, C. Landers, B. Saxton, E. Felix,
B. Gilbert, J. Sederstrom, A. Varadhachary, M. M. Singh, D. Chatterjee, D. B. Corry, and
J. S. McMurray. 2018. “Small molecule targeting of the STAT5/6 Src homology 2 (SH2) domains to
inhibit allergic airway disease.” J Biol Chem 293 (26):10026–40. doi: 10.1074/jbc.RA117.000567.
Kotaja, N., U. Karvonen, O. A. Janne, and J. J. Palvimo. 2002. “PIAS proteins modulate transcription
factors by functioning as SUMO-1 ligases.” Mol Cell Biol 22 (14):5222–34. doi: 10.1128/
mcb.22.14.5222-5234.2002.
Kovarik, P., M. Mangold, K. Ramsauer, H. Heidari, R. Steinborn, A. Zotter, D. E. Levy, M. Muller, and
T. Decker. 2001. “Specificity of signaling by STAT1 depends on SH2 and C-terminal domains that
regulate Ser727 phosphorylation, differentially affecting specific target gene expression.” EMBO J
20 (1–2):91–100. doi: 10.1093/emboj/20.1.91.
Krebs, D. L., and D. J. Hilton. 2001. “SOCS proteins: negative regulators of cytokine signaling.” Stem
Cells 19 (5):378–87. doi: 10.1634/stemcells.19-5-378.
Kurzer, J. H., L. S. Argetsinger, Y. J. Zhou, J. L. Kouadio, J. J. O’Shea, and C. Carter-Su. 2004. “Tyrosine
813 is a site of JAK2 autophosphorylation critical for activation of JAK2 by SH2-B beta.” Mol Cell
Biol 24 (10):4557–70. doi: 10.1128/mcb.24.10.4557-4570.2004.
LaFave, L. M., and R. L. Levine. 2012. “JAK2 the future: therapeutic strategies for JAK-dependent
malignancies.” Trends Pharmacol Sci 33 (11):574–82. doi: 10.1016/j.tips.2012.08.005.
JAK-STAT Structure Function 29

Larner, A. C., M. David, G. M. Feldman, K. Igarashi, R. H. Hackett, D. S. Webb, S. M. Sweitzer,


E. F. Petricoin, 3rd, and D. S. Finbloom. 1993. “Tyrosine phosphorylation of DNA binding
proteins by multiple cytokines.” Science 261 (5129):1730–33. doi: 10.1126/science.8378773.
Lee, B. J., Y. J. Kim, D. H. Cho, N. W. Sohn, and H. Kang. 2011. “Immunomodulatory effect of
water extract of cinnamon on anti-CD3-induced cytokine responses and p38, JNK, ERK1/2,
and STAT4 activation.” Immunopharmacol Immunotoxicol 33 (4):714–22. doi: 10.3109/
08923973.2011.564185.
Lin, L., B. Hutzen, P. K. Li, S. Ball, M. Zuo, S. DeAngelis, E. Foust, M. Sobo, L. Friedman, D. Bhasin,
L. Cen, C. Li, and J. Lin. 2010. “A novel small molecule, LLL12, inhibits STAT3 phosphorylation
and activities and exhibits potent growth-suppressive activity in human cancer cells.” Neoplasia 12
(1):39–50. doi: 10.1593/neo.91196.
Liongue, C., L. A. O’Sullivan, M. C. Trengove, and A. C. Ward. 2012. “Evolution of JAK–STAT
pathway components: mechanisms and role in immune system development.” PLoS One 7 (3):
e32777. doi: 10.1371/journal.pone.0032777.
Liongue, C., and A. C. Ward. 2013. “Evolution of the JAK–STAT pathway.” JAKSTAT 2 (1):e22756.
doi: 10.4161/jkst.22756.
Liu, B., J. Liao, X. Rao, S. A. Kushner, C. D. Chung, D. D. Chang, and K. Shuai. 1998. “Inhibition of
Stat1-mediated gene activation by PIAS1.” Proc Natl Acad Sci U S A 95 (18):10626–31. doi:
10.1073/pnas.95.18.10626.
Liu, B. A., K. Jablonowski, M. Raina, M. Arce, T. Pawson, and P. D. Nash. 2006. “The human and mouse
complement of SH2 domain proteins-establishing the boundaries of phosphotyrosine signaling.”
Mol Cell 22 (6):851–68. doi: 10.1016/j.molcel.2006.06.001.
Liu, F., D. E. Hill, and J. Chernoff. 1996. “Direct binding of the proline-rich region of protein tyrosine
phosphatase 1B to the Src homology 3 domain of p130(Cas).” J Biol Chem 271 (49):31290–95. doi:
10.1074/jbc.271.49.31290.
Liu, L., K. M. McBride, and N. C. Reich. 2005. “STAT3 nuclear import is independent of tyrosine
phosphorylation and mediated by importin-alpha3.” Proc Natl Acad Sci U S A 102 (23):8150–55.
doi: 10.1073/pnas.0501643102.
Lorenz, U. 2009. “SHP-1 and SHP-2 in T cells: two phosphatases functioning at many levels.” Immunol
Rev 228 (1):342–59. doi: 10.1111/j.1600-065X.2008.00760.x.
Lucet, I. S., E. Fantino, M. Styles, R. Bamert, O. Patel, S. E. Broughton, M. Walter, C. J. Burns,
H. Treutlein, A. F. Wilks, and J. Rossjohn. 2006. “The structural basis of Janus kinase 2 inhibition
by a potent and specific pan-Janus kinase inhibitor.” Blood 107 (1):176–83. doi: 10.1182/blood­
2005-06-2413.
Lupardus, P. J., G. Skiniotis, A. J. Rice, C. Thomas, S. Fischer, T. Walz, and K. C. Garcia. 2011.
“Structural snapshots of full-length Jak1, a transmembrane gp130/IL-6/IL-6Ralpha cytokine
receptor complex, and the receptor-Jak1 holocomplex.” Structure 19 (1):45–55. doi: 10.1016/j.
str.2010.10.010.
MacDonagh, L., S. G. Gray, E. Breen, S. Cuffe, S. P. Finn, K. J. O’Byrne, and M. P. Barr. 2018. “BBI608
inhibits cancer stemness and reverses cisplatin resistance in NSCLC.” Cancer Lett 428:117–26. doi:
10.1016/j.canlet.2018.04.008.
Machida, K., C. M. Thompson, K. Dierck, K. Jablonowski, S. Karkkainen, B. Liu, H. Zhang,
P. D. Nash, D. K. Newman, P. Nollau, T. Pawson, G. H. Renkema, K. Saksela, M. R. Schiller,
D. G. Shin, and B. J. Mayer. 2007. “High-throughput phosphotyrosine profiling using SH2
domains.” Mol Cell 26 (6):899–915. doi: 10.1016/j.molcel.2007.05.031.
Manning, G., D. B. Whyte, R. Martinez, T. Hunter, and S. Sudarsanam. 2002. “The protein kinase
complement of the human genome.” Science 298 (5600):1912–34. doi: 10.1126/science.1075762.
Matsuo, K., A. Bettaieb, N. Nagata, I. Matsuo, H. Keilhack, and F. G. Haj. 2011. “Regulation of brown
fat adipogenesis by protein tyrosine phosphatase 1B.” PLoS One 6 (1):e16446. doi: 10.1371/journal.
pone.0016446.
McNally, R., A. V. Toms, and M. J. Eck. 2016. “Crystal structure of the FERM-SH2 module of human
Jak2.” PLoS One 11 (5):e0156218. doi: 10.1371/journal.pone.0156218.
Meng, H., C. Yang, W. Ni, W. Ding, X. Yang, and W. Qian. 2007. “Antitumor activity of fludarabine
against human multiple myeloma in vitro and in vivo.” Eur J Haematol 79 (6):486–93. doi: 10.1111/
j.1600-0609.2007.00968.x.
30 JAK-STAT Signaling in Diseases

Mertens, C., M. Zhong, R. Krishnaraj, W. Zou, X. Chen, and J. E. Darnell, Jr. 2006. “Dephosphorylation
of phosphotyrosine on STAT1 dimers requires extensive spatial reorientation of the monomers
facilitated by the N-terminal domain.” Genes Dev 20 (24):3372–81. doi: 10.1101/gad.1485406.
Metcalf, D., C. J. Greenhalgh, E. Viney, T. A. Willson, R. Starr, N. A. Nicola, D. J. Hilton, and
W. S. Alexander. 2000. “Gigantism in mice lacking suppressor of cytokine signalling-2.” Nature
405 (6790):1069–73. doi: 10.1038/35016611.
Migita, K., T. Miyashita, Y. Izumi, T. Koga, A. Komori, Y. Maeda, Y. Jiuchi, Y. Aiba, S. Yamasaki,
A. Kawakami, M. Nakamura, and H. Ishibashi. 2011. “Inhibitory effects of the JAK inhibitor
CP690,550 on human CD4(+) T lymphocyte cytokine production.” BMC Immunol 12:51. doi:
10.1186/1471-2172-12-51.
Muller, J., B. Sperl, W. Reindl, A. Kiessling, and T. Berg. 2008. “Discovery of chromone-based inhibitors

of the transcription factor STAT5.” Chembiochem 9 (5):723–27. doi: 10.1002/cbic.200700701.

Myers, M. P., J. N. Andersen, A. Cheng, M. L. Tremblay, C. M. Horvath, J. P. Parisien, A. Salmeen,

D. Barford, and N. K. Tonks. 2001. “TYK2 and JAK2 are substrates of protein-tyrosine phospha­
tase 1B.” J Biol Chem 276 (51):47771–74. doi: 10.1074/jbc.C100583200.
Nagashima, S., T. Hondo, H. Nagata, T. Ogiyama, J. Maeda, H. Hoshii, T. Kontani, S. Kuromitsu,
K. Ohga, M. Orita, K. Ohno, A. Moritomo, K. Shiozuka, M. Furutani, M. Takeuchi, M. Ohta, and
S. Tsukamoto. 2009. “Novel 7H-pyrrolo[2,3-d]pyrimidine derivatives as potent and orally active
STAT6 inhibitors.” Bioorg Med Chem 17 (19):6926–36. doi: 10.1016/j.bmc.2009.08.021.
Nagashima, S., M. Yokota, E. Nakai, S. Kuromitsu, K. Ohga, M. Takeuchi, S. Tsukamoto, and M. Ohta.
2007. “Synthesis and evaluation of 2-{[2-(4-hydroxyphenyl)-ethyl]amino}pyrimidine-5-carboxa­
mide derivatives as novel STAT6 inhibitors.” Bioorg Med Chem 15 (2):1044–55. doi: 10.1016/j.
bmc.2006.10.015.
Naka, T., M. Narazaki, M. Hirata, T. Matsumoto, S. Minamoto, A. Aono, N. Nishimoto, T. Kajita,
T. Taga, K. Yoshizaki, S. Akira, and T. Kishimoto. 1997. “Structure and function of a new
STAT-induced STAT inhibitor.” Nature 387 (6636):924–29. doi: 10.1038/43219.
Nelson, E. A., S. R. Walker, E. Weisberg, M. Bar-Natan, R. Barrett, L. B. Gashin, S. Terrell,
J. L. Klitgaard, L. Santo, M. R. Addorio, B. L. Ebert, J. D. Griffin, and D. A. Frank. 2011. “The
STAT5 inhibitor pimozide decreases survival of chronic myelogenous leukemia cells resistant to
kinase inhibitors.” Blood 117 (12):3421–29. doi: 10.1182/blood-2009-11-255232.
Nicholson, S. E., T. A. Willson, A. Farley, R. Starr, J. G. Zhang, M. Baca, W. S. Alexander, D. Metcalf,
D. J. Hilton, and N. A. Nicola. 1999. “Mutational analyses of the SOCS proteins suggest a dual
domain requirement but distinct mechanisms for inhibition of LIF and IL-6 signal transduction.”
EMBO J 18 (2):375–85. doi: 10.1093/emboj/18.2.375.
Niu, G. J., J. D. Xu, W. J. Yuan, J. J. Sun, M. C. Yang, Z. H. He, X. F. Zhao, and J. X. Wang. 2018.
“Protein inhibitor of activated STAT (PIAS) negatively regulates the JAK–STAT pathway by
inhibiting STAT phosphorylation and translocation.” Front Immunol 9:2392. doi: 10.3389/
fimmu.2018.02392.
Nosaka, T., T. Kawashima, K. Misawa, K. Ikuta, A. L. Mui, and T. Kitamura. 1999. “STAT5 as
a molecular regulator of proliferation, differentiation and apoptosis in hematopoietic cells.”
EMBO J 18 (17):4754–65. doi: 10.1093/emboj/18.17.4754.
O’Brien, K. B., J. J. O’Shea, and C. Carter-Su. 2002. “SH2-B family members differentially regulate JAK
family tyrosine kinases.” J Biol Chem 277 (10):8673–81. doi: 10.1074/jbc.M109165200.
Onimoe, G. I., A. Liu, L. Lin, C. C. Wei, E. B. Schwartz, D. Bhasin, C. Li, J. R. Fuchs, P. K. Li,
P. Houghton, A. Termuhlen, T. Gross, and J. Lin. 2012. “Small molecules, LLL12 and FLLL32,
inhibit STAT3 and exhibit potent growth suppressive activity in osteosarcoma cells and tumor
growth in mice.” Invest New Drugs 30 (3):916–26. doi: 10.1007/s10637-011-9645-1.
Pan, Y., F. Zhou, R. Zhang, and F. X. Claret. 2013. “Stat3 inhibitor stattic exhibits potent antitumor
activity and induces chemo- and radio-sensitivity in nasopharyngeal carcinoma.” PLoS One 8 (1):
e54565. doi: 10.1371/journal.pone.0054565.
Pao, L. I., K. Badour, K. A. Siminovitch, and B. G. Neel. 2007. “Nonreceptor protein-tyrosine
phosphatases in immune cell signaling.” Annu Rev Immunol 25:473–523. doi: 10.1146/annurev.
immunol.23.021704.115647.
Park, J. S., S. K. Kwok, M. A. Lim, E. K. Kim, J. G. Ryu, S. M. Kim, H. J. Oh, J. H. Ju, S. H. Park,
H. Y. Kim, and M. L. Cho. 2014. “STA-21, a promising STAT-3 inhibitor that reciprocally regulates
JAK-STAT Structure Function 31

Th17 and Treg cells, inhibits osteoclastogenesis in mice and humans and alleviates autoimmune
inflammation in an experimental model of rheumatoid arthritis.” Arthritis Rheumatol 66 (4):918–29.
doi: 10.1002/art.38305.
Poole, A. W., and M. L. Jones. 2005. “A SHPing tale: perspectives on the regulation of SHP-1 and SHP-2
tyrosine phosphatases by the C-terminal tail.” Cell Signal 17 (11):1323–32. doi: 10.1016/j.
cellsig.2005.05.016.
Rajendrakumar, G. V., V. Radha, and G. Swarup. 1993. “Stabilization of a protein-tyrosine phosphatase
mRNA upon mitogenic stimulation of T-lymphocytes.” Biochim Biophys Acta 1216 (2):205–12. doi:
10.1016/0167-4781(93)90146-5.
Rath, K. S., S. K. Naidu, P. Lata, H. K. Bid, B. K. Rivera, G. A. McCann, B. J. Tierney, A. C. Elnaggar,
V. Bravo, G. Leone, P. Houghton, K. Hideg, P. Kuppusamy, D. E. Cohn, and K. Selvendiran. 2014.
“HO-3867, a safe STAT3 inhibitor, is selectively cytotoxic to ovarian cancer.” Cancer Res 74
(8):2316–27. doi: 10.1158/0008-5472.CAN-13-2433.
Rawlings, J. S., K. M. Rosler, and D. A. Harrison. 2004. “The JAK–STAT signaling pathway.” J Cell Sci
117 (Pt 8):1281–83. doi: 10.1242/jcs.00963.
Reich, N. C. 2013. “STATs get their move on.” JAKSTAT 2 (4):e27080. doi: 10.4161/jkst.27080.
Ren, Y., L. Lu, T. B. Guo, J. Qiu, Y. Yang, A. Liu, and J. Z. Zhang. 2008. “Novel immunomodulatory
properties of berbamine through selective down-regulation of STAT4 and action of IFN-gamma in
experimental autoimmune encephalomyelitis.” J Immunol 181 (2):1491–98. doi: 10.4049/
jimmunol.181.2.1491.
Riehle, K. J., J. S. Campbell, R. S. McMahan, M. M. Johnson, R. P. Beyer, T. K. Bammler, and N. Fausto.
2008. “Regulation of liver regeneration and hepatocarcinogenesis by suppressor of cytokine
signaling 3.” J Exp Med 205 (1):91–103. doi: 10.1084/jem.20070820.
Roth, W., C. Sustmann, M. Kieslinger, A. Gilmozzi, D. Irmer, E. Kremmer, C. Turck, and R. Grosschedl.
2004. “PIASy-deficient mice display modest defects in IFN and Wnt signaling.” J Immunol 173
(10):6189–99. doi: 10.4049/jimmunol.173.10.6189.
Ruff-Jamison, S., K. Chen, and S. Cohen. 1993. “Induction by EGF and interferon-gamma of tyrosine
phosphorylated DNA binding proteins in mouse liver nuclei.” Science 261 (5129):1733–36. doi:
10.1126/science.8378774.
Sachdev, S., L. Bruhn, H. Sieber, A. Pichler, F. Melchior, and R. Grosschedl. 2001. “PIASy, a nuclear
matrix-associated SUMO E3 ligase, represses LEF1 activity by sequestration into nuclear bodies.”
Genes Dev 15 (23):3088–103. doi: 10.1101/gad.944801.
Sadowski, H. B., K. Shuai, J. E. Darnell, Jr., and M. Z. Gilman. 1993. “A common nuclear signal
transduction pathway activated by growth factor and cytokine receptors.” Science 261 (5129):1739–44.
doi: 10.1126/science.8397445.
Saharinen, P., K. Takaluoma, and O. Silvennoinen. 2000. “Regulation of the Jak2 tyrosine kinase
by its pseudokinase domain.” Mol Cell Biol 20 (10):3387–95. doi: 10.1128/mcb.20.10.3387­
3395.2000.
Saharinen, P., M. Vihinen, and O. Silvennoinen. 2003. “Autoinhibition of Jak2 tyrosine kinase is
dependent on specific regions in its pseudokinase domain.” Mol Biol Cell 14 (4):1448–59. doi:
10.1091/mbc.e02-06-0342.
Sakurai, M., M. Nishio, K. Yamamoto, T. Okuda, K. Kawano, and T. Ohnuki. 2003. “TMC-264,
a novel inhibitor of STAT6 activation produced by Phoma sp. TC 1674.” J Antibiot (Tokyo) 56
(6):513–19.
Sasaki, A., H. Yasukawa, A. Suzuki, S. Kamizono, T. Syoda, I. Kinjyo, M. Sasaki, J. A. Johnston, and
A. Yoshimura. 1999. “Cytokine-inducible SH2 protein-3 (CIS3/SOCS3) inhibits Janus tyrosine kinase
by binding through the N-terminal kinase inhibitory region as well as SH2 domain.” Genes Cells 4
(6):339–51.
Schindler, C., and I. Strehlow. 2000. “Cytokines and STAT signaling.” Adv Pharmacol 47:113–74.
Schust, J., B. Sperl, A. Hollis, T. U. Mayer, and T. Berg. 2006. “Stattic: a small-molecule inhibitor
of STAT3 activation and dimerization.” Chem Biol 13 (11):1235–42. doi: 10.1016/j.
chembiol.2006.09.018.
Scully, S. S., A. J. Tang, M. Lundh, C. M. Mosher, K. M. Perkins, and B. K. Wagner. 2013. “Small­
molecule inhibitors of cytokine-mediated STAT1 signal transduction in beta-cells with improved
aqueous solubility.” J Med Chem 56 (10):4125–29. doi: 10.1021/jm400397x.
32 JAK-STAT Signaling in Diseases

Seif, F., M. Khoshmirsafa, H. Aazami, M. Mohsenzadegan, G. Sedighi, and M. Bahar. 2017. “The role of
JAK–STAT signaling pathway and its regulators in the fate of T helper cells.” Cell Commun Signal
15 (1):23. doi: 10.1186/s12964-017-0177-y.
Sekimoto, T., N. Imamoto, K. Nakajima, T. Hirano, and Y. Yoneda. 1997. “Extracellular
signal-dependent nuclear import of Stat1 is mediated by nuclear pore-targeting complex formation
with NPI-1, but not Rch1.” EMBO J 16 (23):7067–77. doi: 10.1093/emboj/16.23.7067.
Selvendiran, K., S. Ahmed, A. Dayton, M. L. Kuppusamy, B. K. Rivera, T. Kalai, K. Hideg, and
P. Kuppusamy. 2011. “HO-3867, a curcumin analog, sensitizes cisplatin-resistant ovarian carci­
noma, leading to therapeutic synergy through STAT3 inhibition.” Cancer Biol Ther 12 (9):837–45.
doi: 10.4161/cbt.12.9.17713.
Shuai, K., C. Schindler, V. R. Prezioso, and J. E. Darnell, Jr. 1992. “Activation of transcription by
IFN-gamma: tyrosine phosphorylation of a 91-kD DNA binding protein.” Science 258
(5089):1808–12. doi: 10.1126/science.1281555.
Siddiquee, K., S. Zhang, W. C. Guida, M. A. Blaskovich, B. Greedy, H. R. Lawrence, M. L. Yip, R. Jove,
M. M. McLaughlin, N. J. Lawrence, S. M. Sebti, and J. Turkson. 2007a. “Selective chemical probe
inhibitor of Stat3, identified through structure-based virtual screening, induces antitumor activity.”
Proc Natl Acad Sci U S A 104 (18):7391–96. doi: 10.1073/pnas.0609757104.
Siddiquee, K. A., P. T. Gunning, M. Glenn, W. P. Katt, S. Zhang, C. Schrock, S. M. Sebti, R. Jove,
A. D. Hamilton, and J. Turkson. 2007b. “An oxazole-based small-molecule Stat3 inhibitor
modulates Stat3 stability and processing and induces antitumor cell effects.” ACS Chem Biol 2
(12):787–98. doi: 10.1021/cb7001973.
Silvennoinen, O., C. Schindler, J. Schlessinger, and D. E. Levy. 1993. “Ras-independent growth factor
signaling by transcription factor tyrosine phosphorylation.” Science 261 (5129):1736–39. doi:
10.1126/science.8378775.
Sommer, U., C. Schmid, R. M. Sobota, U. Lehmann, N. J. Stevenson, J. A. Johnston, F. Schaper,
P. C. Heinrich, and S. Haan. 2005. “Mechanisms of SOCS3 phosphorylation upon interleukin-6
stimulation. Contributions of Src- and receptor-tyrosine kinases.” J Biol Chem 280 (36):31478–88.
doi: 10.1074/jbc.M506008200.
Song, H., R. Wang, S. Wang, and J. Lin. 2005. “A low-molecular-weight compound discovered through
virtual database screening inhibits Stat3 function in breast cancer cells.” Proc Natl Acad Sci U S A
102 (13):4700–05. doi: 10.1073/pnas.0409894102.
Stark, G. R., and J. E. Darnell, Jr. 2012. “The JAK–STAT pathway at twenty.” Immunity 36 (4):503–14.
doi: 10.1016/j.immuni.2012.03.013.
Starr, R., T. A. Willson, E. M. Viney, L. J. Murray, J. R. Rayner, B. J. Jenkins, T. J. Gonda,
W. S. Alexander, D. Metcalf, N. A. Nicola, and D. J. Hilton. 1997. “A family of cytokine-inducible
inhibitors of signalling.” Nature 387 (6636):917–21. doi: 10.1038/43206.
Subramaniam, A., M. K. Shanmugam, E. Perumal, F. Li, A. Nachiyappan, X. Dai, S. N. Swamy,
K. S. Ahn, A. P. Kumar, B. K. Tan, K. M. Hui, and G. Sethi. 2013. “Potential role of signal
transducer and activator of transcription (STAT)3 signaling pathway in inflammation, survival,
proliferation and invasion of hepatocellular carcinoma.” Biochim Biophys Acta 1835 (1):46–60. doi:
10.1016/j.bbcan.2012.10.002.
Subramaniam, P. S., B. A. Torres, and H. M. Johnson. 2001. “So many ligands, so few transcription
factors: a new paradigm for signaling through the STAT transcription factors.” Cytokine 15
(4):175–87. doi: 10.1006/cyto.2001.0905.
Sutherland, K. D., G. J. Lindeman, D. Y. Choong, S. Wittlin, L. Brentzell, W. Phillips, I. G. Campbell,
and J. E. Visvader. 2004. “Differential hypermethylation of SOCS genes in ovarian and breast
carcinomas.” Oncogene 23 (46):7726–33. doi: 10.1038/sj.onc.1207787.
Tanaka, T., M. A. Soriano, and M. J. Grusby. 2005. “SLIM is a nuclear ubiquitin E3 ligase that negatively
regulates STAT signaling.” Immunity 22 (6):729–36. doi: 10.1016/j.immuni.2005.04.008.
Taylor, S. S., D. R. Knighton, J. Zheng, L. F. Ten Eyck, and J. M. Sowadski. 1992. “Structural framework for
the protein kinase family.” Annu Rev Cell Biol 8:429–62. doi: 10.1146/annurev.cb.08.110192.002241.
Tepass, U. 2009. “FERM proteins in animal morphogenesis.” Curr Opin Genet Dev 19 (4):357–67. doi:
10.1016/j.gde.2009.05.006.
Tierney, B. J., G. A. McCann, D. E. Cohn, E. Eisenhauer, M. Sudhakar, P. Kuppusamy, K. Hideg, and
K. Selvendiran. 2012. “HO-3867, a STAT3 inhibitor induces apoptosis by inactivation of STAT3
JAK-STAT Structure Function 33

activity in BRCA1-mutated ovarian cancer cells.” Cancer Biol Ther 13 (9):766–75. doi: 10.4161/
cbt.20559.
Toms, A. V., A. Deshpande, R. McNally, Y. Jeong, J. M. Rogers, C. U. Kim, S. M. Gruner, S. B. Ficarro,
J. A. Marto, M. Sattler, J. D. Griffin, and M. J. Eck. 2013. “Structure of a pseudokinase-domain
switch that controls oncogenic activation of Jak kinases.” Nat Struct Mol Biol 20 (10):1221–23. doi:
10.1038/nsmb.2673.
Tonks, N. K. 2003. “PTP1B: from the sidelines to the front lines!” FEBS Lett 546 (1):140–48. doi:
10.1016/s0014-5793(03)00603-3.
Tonks, N. K. 2006. “Protein tyrosine phosphatases: from genes, to function, to disease.” Nat Rev Mol Cell
Biol 7 (11):833–46. doi: 10.1038/nrm2039.
Tonks, N. K., C. D. Diltz, and E. H. Fischer. 1988. “Purification of the major protein-tyrosine­
phosphatases of human placenta.” J Biol Chem 263 (14):6722–30.
Turkson, J., J. S. Kim, S. Zhang, J. Yuan, M. Huang, M. Glenn, E. Haura, S. Sebti, A. D. Hamilton, and
R. Jove. 2004. “Novel peptidomimetic inhibitors of signal transducer and activator of transcription
3 dimerization and biological activity.” Mol Cancer Ther 3 (3):261–69.
Ungureanu, D., J. Wu, T. Pekkala, Y. Niranjan, C. Young, O. N. Jensen, C. F. Xu, T. A. Neubert,
R. C. Skoda, S. R. Hubbard, and O. Silvennoinen. 2011. “The pseudokinase domain of JAK2 is a
dual-specificity protein kinase that negatively regulates cytokine signaling.” Nat Struct Mol Biol 18
(9):971–76. doi: 10.1038/nsmb.2099.
Vogelstein, B., N. Papadopoulos, V. E. Velculescu, S. Zhou, L. A. Diaz, Jr., and K. W. Kinzler. 2013.
“Cancer genome landscapes.” Science 339 (6127):1546–58. doi: 10.1126/science.1235122.
Wagner, M. J., M. M. Stacey, B. A. Liu, and T. Pawson. 2013. “Molecular mechanisms of SH2- and
PTB-domain-containing proteins in receptor tyrosine kinase signaling.” Cold Spring Harb Perspect
Biol 5 (12):a008987. doi: 10.1101/cshperspect.a008987.
Wallweber, H. J., C. Tam, Y. Franke, M. A. Starovasnik, and P. J. Lupardus. 2014. “Structural basis of
recognition of interferon-alpha receptor by tyrosine kinase 2.” Nat Struct Mol Biol 21 (5):443–48.
doi: 10.1038/nsmb.2807.
Wang, Y., S. Wang, Y. Wu, Y. Ren, Z. Li, X. Yao, C. Zhang, N. Ye, C. Jing, J. Dong, K. Zhang, S. Sun,
M. Zhao, W. Guo, X. Qu, Y. Qiao, H. Chen, L. Kong, R. Jin, X. Wang, L. Zhang, J. Zhou, Q. Shen,
and X. Zhou. 2017. “Suppression of the growth and invasion of human head and neck squamous
cell carcinomas via regulating STAT3 signaling and the miR-21/beta-catenin axis with HJC0152.”
Mol Cancer Ther 16 (4):578–90. doi: 10.1158/1535-7163.MCT-16-0606.
Wang, Z., J. Li, W. Xiao, J. Long, and H. Zhang. 2018. “The STAT3 inhibitor S3I-201 suppresses fibrogenesis
and angiogenesis in liver fibrosis.” Lab Invest 98 (12):1600–13. doi: 10.1038/s41374-018-0127-3.
Wilks, A. F. 1989. “Two putative protein-tyrosine kinases identified by application of the polymerase
chain reaction.” Proc Natl Acad Sci U S A 86 (5):1603–07. doi: 10.1073/pnas. 86.5.1603.
Wilks, A. F. 2008. “The JAK kinases: not just another kinase drug discovery target.” Semin Cell Dev Biol
19 (4):319–28. doi: 10.1016/j.semcdb.2008.07.020.
Williams, L., L. Bradley, A. Smith, and B. Foxwell. 2004. “Signal transducer and activator of transcrip­
tion 3 is the dominant mediator of the anti-inflammatory effects of IL-10 in human macrophages.”
J Immunol 172 (1):567–76. doi: 10.4049/jimmunol.172.1.567.
Wingelhofer, B., B. Maurer, E. C. Heyes, A. A. Cumaraswamy, A. Berger-Becvar, E. D. de Araujo,
A. Orlova, P. Freund, F. Ruge, J. Park, G. Tin, S. Ahmar, C. H. Lardeau, I. Sadovnik, D. Bajusz,
G. M. Keseru, F. Grebien, S. Kubicek, P. Valent, P. T. Gunning, and R. Moriggl. 2018. “Pharma­
cologic inhibition of STAT5 in acute myeloid leukemia.” Leukemia 32 (5):1135–46. doi: 10.1038/
s41375-017-0005-9.
Wolf, C., and M. R. Mofrad. 2008. “On the octagonal structure of the nuclear pore complex: insights
from coarse-grained models.” Biophys J 95 (4):2073–85. doi: 10.1529/biophysj.108.130336.
Wu, C., M. Sun, L. Liu, and G. W. Zhou. 2003. “The function of the protein tyrosine phosphatase SHP-1
in cancer.” Gene 306:1–12.
Xu, X., Y. L. Sun, and T. Hoey. 1996. “Cooperative DNA binding and sequence-selective recognition
conferred by the STAT amino-terminal domain.” Science 273 (5276):794–97. doi: 10.1126/
science.273.5276.794.
Yamaoka, K., P. Saharinen, M. Pesu, V. E. Holt, 3rd, O. Silvennoinen, and J. J. O’Shea. 2004. “The Janus
kinases (Jaks).” Genome Biol 5 (12):253. doi: 10.1186/gb-2004-5-12-253.
34 JAK-STAT Signaling in Diseases

Yasukawa, H., H. Misawa, H. Sakamoto, M. Masuhara, A. Sasaki, T. Wakioka, S. Ohtsuka,


T. Imaizumi, T. Matsuda, J. N. Ihle, and A. Yoshimura. 1999. “The JAK-binding protein JAB
inhibits Janus tyrosine kinase activity through binding in the activation loop.” EMBO J 18
(5):1309–20. doi: 10.1093/emboj/18.5.1309.
Yoshimura, A., and H. Yasukawa. 2012. “JAK’s SOCS: a mechanism of inhibition.” Immunity 36
(2):157–59. doi: 10.1016/j.immuni.2012.01.010.
You-Ten, K. E., E. S. Muise, A. Itie, E. Michaliszyn, J. Wagner, S. Jothy, W. S. Lapp, and M. L. Tremblay.
1997. “Impaired bone marrow microenvironment and immune function in T cell protein tyrosine
phosphatase-deficient mice.” J Exp Med 186 (5):683–93. doi: 10.1084/jem.186.5.683.
Yu, H., and R. Jove. 2004. “The STATs of cancer–new molecular targets come of age.” Nat Rev Cancer 4
(2):97–105. doi: 10.1038/nrc1275.
Zhang, D., A. Wlodawer, and J. Lubkowski. 2016. “Crystal structure of a complex of the intracellular
domain of interferon lambda receptor 1 (IFNLR1) and the FERM/SH2 domains of human JAK1.”
J Mol Biol 428 (23):4651–68. doi: 10.1016/j.jmb.2016.10.005.
Zhang, J. G., A. Farley, S. E. Nicholson, T. A. Willson, L. M. Zugaro, R. J. Simpson, R. L. Moritz,
D. Cary, R. Richardson, G. Hausmann, B. T. Kile, S. B. Kent, W. S. Alexander, D. Metcalf,
D. J. Hilton, N. A. Nicola, and M. Baca. 1999. “The conserved SOCS box motif in suppressors of
cytokine signaling binds to elongins B and C and may couple bound proteins to proteasomal
degradation.” Proc Natl Acad Sci U S A 96 (5):2071–76. doi: 10.1073/pnas.96.5.2071.
Zhang, X., Y. Sun, R. Pireddu, H. Yang, M. K. Urlam, H. R. Lawrence, W. C. Guida, N. J. Lawrence, and
S. M. Sebti. 2013. “A novel inhibitor of STAT3 homodimerization selectively suppresses STAT3
activity and malignant transformation.” Cancer Res 73 (6):1922–33. doi: 10.1158/0008-5472.CAN­
12-3175.
Zhang, X., P. Yue, S. Fletcher, W. Zhao, P. T. Gunning, and J. Turkson. 2010. “A novel small-molecule
disrupts Stat3 SH2 domain-phosphotyrosine interactions and Stat3-dependent tumor processes.”
Biochem Pharmacol 79 (10):1398–409. doi: 10.1016/j.bcp.2010.01.001.
Zhang, X., P. Yue, B. D. Page, T. Li, W. Zhao, A. T. Namanja, D. Paladino, J. Zhao, Y. Chen,
P. T. Gunning, and J. Turkson. 2012. “Orally bioavailable small-molecule inhibitor of transcription
factor Stat3 regresses human breast and lung cancer xenografts.” Proc Natl Acad Sci U S A 109
(24):9623–28. doi: 10.1073/pnas.1121606109.
Zhang, Y., Z. Jin, H. Zhou, X. Ou, Y. Xu, H. Li, C. Liu, and B. Li. 2016. “Suppression of prostate cancer
progression by cancer cell stemness inhibitor napabucasin.” Cancer Med 5 (6):1251–58. doi:
10.1002/cam4.675.
Zhang, Z., K. Shen, W. Lu, and P. A. Cole. 2003. “The role of C-terminal tyrosine phosphorylation in the
regulation of SHP-1 explored via expressed protein ligation.” J Biol Chem 278 (7):4668–74. doi:
10.1074/jbc.M210028200.
Zhao, C., W. Wang, W. Yu, D. Jou, Y. Wang, H. Ma, H. Xiao, H. Qin, C. Zhang, J. Lu, S. Li, C. Li, J. Lin,
and L. Lin. 2016. “A novel small molecule STAT3 inhibitor, LY5, inhibits cell viability, colony
formation, and migration of colon and liver cancer cells.” Oncotarget 7 (11):12917–26. doi:
10.18632/oncotarget.7338.
Zhao, L., H. Dong, C. C. Zhang, L. Kinch, M. Osawa, M. Iacovino, N. V. Grishin, M. Kyba, and
L. J. Huang. 2009. “A JAK2 interdomain linker relays Epo receptor engagement signals to kinase
activation.” J Biol Chem 284 (39):26988–98. doi: 10.1074/jbc.M109.011387.
Zhong, Z., Z. Wen, and J. E. Darnell, Jr. 1994a. “Stat3 and Stat4: members of the family of signal
transducers and activators of transcription.” Proc Natl Acad Sci U S A 91 (11):4806–10. doi:
10.1073/pnas.91.11.4806.
Zhong, Z., Z. Wen, and J. E. Darnell, Jr. 1994b. “Stat3: a STAT family member activated by tyrosine
phosphorylation in response to epidermal growth factor and interleukin-6.” Science 264
(5155):95–98. doi: 10.1126/science.8140422.
Zhou, X. X., P. J. Gao, and B. G. Sun. 2009. “Pravastatin attenuates interferon-gamma action via
modulation of STAT1 to prevent aortic atherosclerosis in apolipoprotein E-knockout mice.” Clin
Exp Pharmacol Physiol 36 (4):373–79. doi: 10.1111/j.1440-1681.2008.05067.x.
3
MicroRNA-Mediated Regulation of JAK-STAT
Signaling in Non-Cancerous Human Diseases

Chandra S. Boosani and Devendra K. Agrawal


Department of Translational Research
College of Osteopathic Medicine of the Pacific
Western University of Health Sciences
Pomona, CA

Wanlin Jiang
Cardiovascular Research Center
Massachusetts General Hospital
Boston MA

Taylor Burke
Department of Clinical and Translational Science
Creighton University School of Medicine
Omaha, NE

3.1 Brief Background


Targeting a signaling pathway that is pathologically active and specific to a given disease could be more
enticing to pursue development of drug targets for disease prevention. Even more is when the same
pathway participates in multiple diseases and plays an important role in progression of the disease.
Certainly, of that kind would be the Janus kinase (JAK)-signal transducers and activators of transcrip­
tion (STAT) signaling pathway that has been identified as an important activator of intra-cellular
signaling molecules, which are associated with many pathological conditions. Notably, this signaling
mechanism promoting key cellular activities was identified to be directly associated with growth,
proliferation, and migration in different cell types. Its role in proliferative disorders, cancer, and
immune regulation was seen to be under tight control by STAT, which is activated by JAK. As
a result of STAT-induced transcriptional activity, expression of several key inflammatory molecules was
reported, which promote the pathological progression of the disease. Both direct and indirect activation
of STAT by JAK has been reported, which denotes their robust involvement and regulation of multiple
cross-connecting pathways, which underscores their participation in diverse cellular mechanisms.
Adding to the already-existing complexity in gene-regulatory mechanisms, the discovery of micro-
RNAs and their proven role in controlling gene expression, layers another dimension of multiple gene
regulations that are driven by a single miRNA (microRNA). At the same time, identification of
miRNA targets also facilitated the predictions of possible cross talks between different pathways. The
miRNAs are short oligomers with full amenability to modify and synthesize. In addition, the ease of
their delivery into the cells and in live animals make them a more favorable choice to use in clinical
research. Further, miRNAs as such have become a tool to drive or prevent their target pathways,
which highlights the potential of miRNAs as another therapeutic avenue to treat human and animal

35
36 JAK-STAT Signaling in Diseases

diseases. Recent advancements using bioinformatic tools have enabled identification of specific loci for
different miRNAs on the chromosomes, which furthered our understanding on the origin and
putative source of miRNAs. The increasing number of miRNAs being identified as potential
therapeutic candidates and their proven success in pre-clinical trials and animal trials show the
promise and safety of miRNAs for disease treatment.
Ample literature evidence suggest that several miRNAs regulate JAK-STAT pathways in cancer,
and many previous publications have previously discussed about this topic. In this chapter, we
specifically tried to exclude discussing about the role of miRNAs that regulate JAK-STAT pathways
in cancer due to the vast abundance and volume of literature available on this topic. At the same
time, very little literature exists that show the role of miRNAs in regulating JAK-STAT pathway in
other diseases (Table 3.1). Therefore, in this chapter, we remain focused on those miRNAs that were
validated to be regulating JAK-STAT pathway in non-cancerous diseases. Although efforts were
made to provide comprehensive information on this topic, the readers are encouraged to refer to the
bibliography below and the current literature for more details.

TABLE 3.1
List of miRNAs and their Target Transcripts that Affect JAK-STAT Signaling in Non-Cancerous Human
Diseases
Anticipated effect
Target of JAK-STAT
miRNA transcript pathway Test model Disease model Reference

miR-19a SOCS1, Inducing effect HuH-7 Liver diseases (Collins et al. 2013, e69090)
SOCS3, immortal cells
SOCS5
miR-155 STAT1 Inhibitory effect Live mice Liver injury (Lv et al. 2015, 6013–6018)
miR-1264 DNMT1 Inhibitory effect Smooth Coronary artery (Boosani et al. 2015,
muscle cells disease 1365–1376)
miR-181a To be char- Inducing effect Dendritic cells Heart disease (Zhu et al. 2017, 2884–2895)
acterized
miR-150 To be char- Inhibitory effect Dendritic cells Heart disease (Zhu et al. 2017, 2884–2895)
acterized
miR-146b STAT5A Inhibitory effect Heart tissue Cardiac (Feng et al. 2014, 361–368)
hypertrophy
miR-499 STAT2 Inhibitory effect Heart tissue Cardiac (Feng et al. 2014, 361–368)
hypertrophy
miR-19a TNF-α, Inducing effect Human Arthritis (Li et al. 2016 2531–2536)
STAT3, PBMCs
SOCS3
miR-21 TNF-α, Inducing effect Human Arthritis (Li et al. 2016 2531–2536)
STAT3, PBMCs
SOCS3
miR-1246 HNFγ Inducing effect Chondrocytes Arthritis (Wu et al. 2017, 2010–2021)
miR-1246 PRKAR1A, Inducing effect Epithelial cells Inflammation and (Bott et al. 2017,
PPP2CB cancer 43897–43914)
miR-203 MCL-1 Inhibitory effect Human carti- Osteoarthritis (Zhao et al. 2017 171–178)
lage cells
miR-29b To be char- Inducing effect Osteoblasts Rheumatoid (Figueiredo Neto and
acterized arthritis Figueiredo 2017, 6365857)
miR-21 To be char- Inducing effect Osteoblasts Rheumatoid (Figueiredo Neto and
acterized arthritis Figueiredo 2017, 6365857)
miR-155 SOCS1 Inducing effect CD4+ T-cells Immune disorders (Yao et al. 2012, e46082)

(Continued )
JAK-STAT Pathway Regulation 37

TABLE 3.1 (Cont.)


Anticipated effect
Target of JAK-STAT
miRNA transcript pathway Test model Disease model Reference

miR-1225-3p GAB3 Inhibitory effect Huh7 cells Influenza A virus (Cheng et al. 2018,
5975–5986)
miR-203 DR1 Promotes JAK­ A549 Influenza A virus (Zhang et al. 2018, 6797­
STAT signaling 018-25073-9)
miR-324-5p CUEDC2 Negative regulator A549 H5N1 (Kumar et al. 2018, 10.1128/
JVI.01057–18. Print 2018
Oct 1)
miR-130a To be char- Inducing effect J6/JFH-1 Hepatitis C viral (Li et al. 2014, 121–128;
acterized immortal cells infection Murayama et al. 2007,
8030–8040)
miR-122 To be char- To be Liver cells Hepatitis C viral (Jopling et al. 2005,
acterized characterized infection 1577–1581)
miR-373 JAK1, IRF9 Inhibitory effect Human liver Hepatitis C viral (Mukherjee et al. 2015,
biopsies infection 3356–3365)
miR-155 SOCS1 Inducing effect HepG2 HBV infection (Lu et al. 2008,
cells, and 10436–10443)
B lymphocytes
miR-103 CDK5R1 Inhibitory effect Stem cells Spinal cord injury (Li et al. 2018, 292–300)
miR-210 Presumably Inhibitory effect In vivo in mice Spinal cord injury (Dai et al. 2018b,
JAK2 6609–6615)
miR-125b JAK1 Inhibitory effect In vivo in mice Axon regeneration/ (Dai et al. 2018a, 582–589)
STAT1 spinal cord injury
miR-155 To be char- Inducing effect Retinal pig- Eye diseases (Kutty et al. 2010, 390–395)
acterized ment epithe­
lial cells
miR-155 To be char- Inducing effect Lymphatic Angiogenesis (Yee et al. 2017,
acterized endothelial 20,683–20,693)
cells
miR-9 SOCS5 Inducing effect Endothelial Angiogenesis (Zhuang et al. 2012,
cells 3513–3523)
miR-155 SOCS1 Inducing effect Microglial Neurodegenerative (Cardoso et al. 2012, 73–88)
cells diseases
miR-146a IL-6 Inducing effect Microglial Neurodegenerative (Saba et al. 2012, e30832)
cells diseases
miR-216a JAK2 Inhibitory effects Neuronal cells Neuroprotection (Tian et al. 2018, 977–988)

3.2 JAK-STAT Targeting miRNAs in Liver Diseases


A group of tumor-suppressor proteins, which together were classified as the suppressor of cytokine
signaling proteins (SOCS), have been well characterized with defined functions toward inhibition of
JAK kinase and STAT transcription factor in the canonical JAK-STAT signaling pathway. The
inhibition of both JAK and STAT by SOCS proteins was found to be mediated through protein
interactions and through epigenetic mechanisms, which are regulated by specific miRNAs. The miRNA
miR-19a targets most of the SOCS transcripts, which include SOCS1, SOCS3, and SOCS5, in addition
to Cullin (Collins et al. 2013, e69090). As JAK-STAT regulation is influenced by SOCS proteins and
since miR-19a inhibits SOCS3, the functional role of miR-19a would be to curtail the inhibitory effects
of SOCS3 on JAK and STAT activities. This was reported recently in immortal “HuH-7 cells”
where IL-6 and IFN-α enhanced activation of pSTAT3 in presence of miR-19a. These results suggest
38 JAK-STAT Signaling in Diseases

that miR-19a indirectly promotes JAK-STAT-mediated expression of pro-inflammatory mediators


(Collins et al. 2013, e69090). It would be more interesting to see if this microRNA miR-19a had any
impact on activation of JAK proteins, since SOCS3 binds to the receptor of JAK in its transmembrane
region and stereotypically prevents its interactions with JAK.
In a significantly large animal study using BALB/c mice with n = 40 in each group, an in vivo study
was performed to evaluate the role of specific miRNAs associated with lipopolysaccharides-(LPS-)
induced sepsis as liver injury model (Lv et al. 2015, 6013–6018). While LPS was reported to have
effectively induced the anticipated liver injury in mice, there was a concomitant increase in the
expression of the miRNA miR-155 in the group that was treated with LPS alone. In the mice group
that were treated with LPS and with miR-155 specific inhibitor, SOCS1 protein was reported to be
significantly upregulated. Further, in presence of the miR-155 specific inhibitor, the expression levels of
STAT1 were shown to be greatly reduced. Also, the levels of tumor necrosis factor alpha (TNF-α) and
interleukin-10 (IL-10) cytokines were reported to be significantly decreased in the mice group that was
treated with LPS and miR-155 specific inhibitor. Since STAT1 drives the expression of the miRNA
miR-155, and miR-155 targets SOCS1 and prevents its expression, the above study provides a clear
mechanistic regulation between the microRNA miR-155 and SOCS1, where miR-155 inhibits SOCS1
to promote JAK-STAT signaling in LPS-induced liver damage (Kutty et al. 2010, 390–395).
Compared to the normal liver tissues and the para-tumor tissues in patients with hepatocellular
carcinoma (HCC), the levels of microRNA miR-409 were reported to be drastically reduced in the
tumor tissues. This downregulation of miR-409 was presumed as a possible cause associated with the
progression of the HCC (Zhang et al. 2019, 146–154). The authors also reported the existence of an
inverse correlation between the expression of miR-409 and the protein levels of JAK2 and STAT3.
Further, in cell-culture system, when HCC cells were transfected with both the microRNA mimic and
inhibitor separately, reduced cell viability and increased apoptosis was observed in the miR-409
transfected cells. As can be anticipated, the opposite results were found in cells that were transfected
with the inhibitor, which blocks the functions of miR-409. Both the cell culture data and the histological
observations indicate that miR-409 targets JAK-STAT pathway to prevent the progression of HCC.
The reports, which showed E74-like factor 2 as a direct target of miR-409 and c-Met as a target for
miR-409-3p, suggest that miR-409 may not be directly targeting JAK or STAT transcripts.
However, all the current evidence point to the same common observation, which indicate that miR­
409 regulates JAK-STAT pathway, and which is more likely to be an indirect cause. In another study,
HCC development was found to be directly correlated with the expression of two miRNAs belonging to
the same class—miR-196a and miR-196b—which regulate JAK-STAT pathway (Ren et al. 2019, 333­
019-1530-4). These two miRNAs were reported to increase cell proliferation, migration, invasion, and
apoptosis induction. However, each of these was found to target two different transcription factors. The
microRNA, miR-196a, primary target was identified as FOXO1, while the target for miR-196b was
FOXP2. Interestingly, both these miRNAs were also found to prevent SOCS2 expression, affecting the
JAK2–STAT5 pathway in HCC. The above reports suggests that in normal hepatic cells these miRNAs
establish a tight control of JAK-STAT pathway as a means to prevent oncogenic transformation.

3.3 Specific MicroRNAs that Regulate JAK-STAT Pathways in Heart Diseases


An indirect epigenetic mechanism, which we recently reported, was the direct targeting of the
main DNA methylase (DNMT1, DNA methyltransferase 1) by miRNA miR-1264 (Boosani et al.
2015, 1365–1376). DNMT1 promotes methylation of CpG residues in the genome DNA, and this
non-specific methylation of DNA would affect all genes, which have CpG islands. Since SOCS3
genomic sequence harbors a well-defined CpG island in its promoter region, it is highly subjected
to methylation by DNMT1. This we recently reported in vascular smooth muscle cells wherein the
miRNA miR-1264, which has strong affinity to bind to the DNMT1 transcript and inhibit its
expression, and thus in absence of DNMT1, SOCS3 expression remain unaffected (Boosani et al.
2015, 1365–1376). Subsequently, SOCS3 was able to prevent the activation of JAK. The miRNA
JAK-STAT Pathway Regulation 39

miR-1264 also prevented the DNMT1 activity induced by TNF-α and insulin-like growth factor 1
(IGF-1) when treated together in smooth muscle cells. These observations establish the existence of
an epigenetic mechanism, which involves both miRNA and a DNA methylase. Also, inhibition of
SOCS3 was identified to be an important cause for the progression of intimal hyperplasia and
restenosis (Gupta et al. 2011, 346–352). Cumulatively, these studies clearly indicate that multiple
mechanisms are involved in preventing SOCS3 expression during restenosis and the miRNA miR­
1264 is a key mediator that can help restore SOCS3 expression to prevent neointima formation in
the arteries.
The role of miRNAs specific to JAK-STAT pathway and its implications in immune regulations
associated with myocardial infarction were reported recently using bone-marrow–derived dendritic
cells from mice (Zhu et al. 2017, 2884–2895). In an interesting approach, the authors used the
cultured supernatant from the necrotic cardiomyocytes as a means to mimic myocardial infarction
environment on dendritic cells. The authors subsequently identified specific miRNAs that can
potentially regulate inflammatory responses mediated by the dendritic cells. The supernatant from
the necrotic cells were reported to upregulate maturation markers in the dendritic cells, such as CD40,
CD83, and CD86, along with increased levels of proinflammatory cytokines. Concurrent increase in
the expression of the miRNA, miR-181a, and simultaneous downregulation of miR-150 was also
reported. Treatment with the necrotic supernatant was shown to activate the JAK-STAT pathway and
also facilitated nuclear translocation of nuclear factor kappa B (NF-κB) and c-Fos. These reports
highlight the significance of the miRNAs—miR-181a and miR-150—in regulating JAK-STAT
signalling, which controls the hypoxic cell death in cardiomyocytes.
In a rat model to study the role of different miRNAs in cardiac hypertrophy and heart failure,
transverse aortic constriction (TAC) was performed in 20 rats (Feng et al. 2014, 361–368). RNA isolated
from the hearts of these rats showed differential expression of multiple miRNAs that regulated different
cellular pathways. The miRNAs, which directly or indirectly targeted JAK-STAT pathway in the above
study, were discussed here. Expression of the microRNAs let-7c, let-7f, miR-1, miR-300-5p, miR-347,
and miR-874 were found downregulated. These miRNAs indirectly act in regulating JAK-STAT
pathway by targeting different interleukins. At the same time, the expression of miRNAs let-7i, miR­
22, miR-29, miR-30a, miR-101, miR-125b-5p, miR-146b, miR-214, and miR-499 were found to be
increased. Notably, only miR-146b and miR-499 were found to act directly on STAT5 and STAT2,
respectively. Cumulatively, this comprehensive miRNA profiling in the cardiac hypertrophy model
resulting from pressure overload due to TAC, identified signature miRNA species. Although a significant
role of JAK-STAT signaling was evident in this study, several other effective target transcripts and other
potent miRNAs regulating different cellular pathways exist which were well characterized.

3.4 MicroRNAs Targeting JAK-STAT Pathways in Arthritis


In efforts to identify specific miRNAs that regulate JAK-STAT signaling in patients with arthritis,
a recent human clinical study was conducted using peripheral blood monocytes (PBMCs) collected
from 20 normal individuals and 20 patients who were suffering from symptomatic juvenile idiopathic
arthritis (Li et al. 2016, 2531–2536). Although the study showed minimal analysis to identify specific
miRNAs, it was interesting to note that two selected miRNAs, miR-19a and miR-21, were reported to
be expressing at a very low level in the patients. Importantly, the authors conclude that TNF-α,
STAT3, and SOCS3 transcripts are the targets for these two miRNAs, miR-19a and miR-21. The
increased expression of TNF-α and STAT3 in the arthritic patients agrees with the other reports in the
literature. However, the expression of SOCS3 can be debatable but it should be noted that SOCS3 has
also been shown to encourage inflammatory responses in a few other diseases, such as in asthma
(Zafra et al. 2014, e91996). Also, the study includes juvenile patients, and presumably is focused on
the local ethnic group, which greatly limits extension of the findings to other population groups unless
evaluated. Additional analysis in this study could have been more beneficial, such as conducting
a microarray-based analysis of all the samples and evaluating the samples to identify what specific
40 JAK-STAT Signaling in Diseases

JAK or STAT molecules are activated through phosphorylation. Another notable limitation in this
study was using real-time PCR as means for quantifying the gene expression. The procedure as such
depends on the primer choice, and thus the estimated gene expression based on the available RNA
needs to support through additional means of evaluation, such as using a protein array or ELISA or
western blotting.
LPS is widely known to initiate many proinflammatory mechanisms in different cell types. Several
volumes of information in the literature shows that inflammatory mediators, which include interleukins,
cytokines, and growth factors, can be upregulated by LPS, both in vitro and in animal models. TNF-α is
certainly one of them, which has the potential to mediate multiple signaling mechanisms and the
evidence is accumulating, which supports the role of TNF-α in inducing epigenetic mechanisms that
promote disease progression. We have earlier identified a clear role of TNF-α in exacerbating
neointimal hyperplasia in the coronary arteries (Gupta et al. 2011, 346–352). Recently, we have also
shown that TNF-α regulates different intracellular pathways including the JAK-STAT signaling (Dhar
et al. 2013, H776-85; Boosani et al. 2015, 1365–1376). LPS-induced expression of IL-1b, IL-6, IL-8, and
TNF-α was identified to promote inflammation in the chondrogenic cells “ATDC5.” LPS-induced
expression of these inflammatory mediators were found to regulate cell viability and apoptotic
pathways that are mediated by JAK1, STAT1, STAT3, Akt, and PI3K (Wu et al. 2017, 2010–2021).
Notably, the regulation of JAK-STAT pathway and its cross-talk with PI3K-Akt pathway by LPS was
found to be mediated through the miRNA, miR-1246. The hepatic nuclear factor gamma “HNFγ”
was reported as a specific target of miR-1246 in the putative chondrocytes. The study highlights
multiple functions of HNFγ, and its inhibition by the miRNA miR-1246 in presence of LPS as
a potential mechanism by which JAK-STAT promotes intracellular inflammation.
In mesenchymal stem cells, the miRNA miR-1246 was reported to be expressed in higher levels
and interestingly, this miRNA expression was independent of the inflammatory signal induced by
TNF-α. However, the main functions of miR-1246 reported was to target and prevent the
expression of two tumor-suppressor genes—cAMP-dependent protein kinase type I-alpha
“PRKAR1A” and the protein phosphates 2C-beta “PPP2CB” (Bott et al. 2017, 43897–43914).
An interesting aspect was that the conditioned medium from the miR-1246 transfected cells was
found to induce JAK-STAT signaling in the normal and tumor epithelial cells.
To understand the role of JAK-STAT pathway and miRNAs that affect JAK-STAT signaling in
osteoarthritis, LPS was used as means to induce inflammation and cellular injury in cultured human
cartilage cells C28/I2 (Zhao et al. 2017, 171–178). The authors report that when chondrocytes were
treated with LPS, expression of a specific miRNA miR-203 was significantly upregulated with
concomitant decrease in cell viability and increase in cellular apoptosis. Further, the authors also
report that overexpression of miR-203 in the chondrocytes exaggerates LPS-induced inflammatory
responses in the transfected cells. Since the direct target of the miRNA, miR-203, is the myeloid
cell leukemia-1 transcript (MCL-1), it was speculated that MCL-1 would play a critical role in
LPS-mediated inflammation during osteoarthritis. Further, as the effector pathways for MCL-1
are Wnt/β-catenin and JAK-STAT, it was speculated that the inflammatory responses initiated
during the progression of osteoarthritis are mediated through upregulation of the miRNA
miR-203, which indirectly affects JAK-STAT pathway through the expression of MCL-1.
In treating rheumatoid arthritis patients, it is not just sufficient to address inflammation since bone
erosion occurs which is also a complication seen in these patients. In this direction, IL-27 has been
shown to promote mineralization in the osteoblasts, which can prevent bone erosion. In attempts to
identify the factors that enhance the effects of IL-27, a microarray-based study revealed that two
specific miRNAs—miR-29b and miR-21—augment the expression of IL-27 in osteoblasts (Figueiredo
Neto and Figueiredo 2017, 6365857). The study showed that treatment with IL-27 not only promoted
osteoblast differentiation, but it also inhibited the proteins that prevent osteoblast formation. Two main
signaling pathways were identified to be associated with osteoblast differentiation, which include JAK­
STAT and TGF-β/BMP/SMAD signaling. In another study using human bone marrow stromal cells
which were induced for osteogenic differentiation, specific miRNAs were identified that were character­
ized as osteogenic specific miRNAs, and these were reported to be targeting JAK-STAT pathway
(Vimalraj and Selvamurugan 2014, 194–202).
JAK-STAT Pathway Regulation 41

3.5 JAK-STAT Specific miRNAs Regulating Immune Responses


CD4+ T-cells are an important class of immune cells with multiple functions and are primarily
involved in suppressing immune reaction. IL-6 and STAT3 were shown to be essentially required
for the differentiation of immune cells, such as T-regulatory and Th17, cells from uncommitted
CD4+ T-cells. In a recent report, inhibition of SOCS1 by the miRNAs miR-155 was shown to
induce differentiation of T-regulatory cells and Th17 cells (Yao et al. 2012, e46082). This cell
differentiation was further supported by the observations that when CD4+ T-cells were trans­
fected with the miRNAs miR-155, increased expression of activated STAT3 and STAT5 were
observed, which correlated with decrease in the expression of SOCS1 in the same group. These
reports clarify a clear role of JAK-STAT signaling in immune cell differentiation, and SOCS1
prevents activation of JAK kinases, which is required to induce STAT3 and STAT5 activation. In
another study to identify the significance and role of miRNAs in regulating proliferation and
death of T-cells, a comprehensive microarray analysis was carried out which was based on the
observations that a drastic decrease in JAK1 expression was noted in IL-2 depleted cells (Ranji
et al. 2015, 169–183). The study identified several miRNAs that can specifically target important
pathway mediators of JAK-STAT signaling to regulate immune responses.

3.6 miRNAs Regulating Viral Diseases Induced by JAK-STAT Pathways


Chronic hepatitis resulting from hepatitis C viral infections can be a major cause that can lead to
hepatocellular carcinoma and other complications in humans. A recent report showed a strong
correlation between the expression of the miRNA miR-130a and the replication of hepatitis C virus
(Li et al. 2014, 121–128). Specifically, the expression of miRNA miR-130a was reported to
significantly inhibit the virus replication in chimeric J6/JFH-1 cells. It is to be noted that the J6/
JFH-1 cell line, which has the chimeric genomic regions that encode J6 structural proteins and
JFH-1 nonstructural proteins, enables the J6/JFH-1 cells to replicate autonomously and produce
infectious HCV particles (Murayama et al. 2007, 8030–8040). The finding that overexpression of
miR-130a in these cells upregulates the expression of type I interferons and other essential proteins
which initiate innate responses clearly indicates the role of interferon in stimulating genes that have
been earlier reported to activate JAK-STAT pathway (Martensen and Justesen 2004, 1–19).
Together, the above reports would provide strong inference, which indicates the potential role of
miR-130a in inhibiting replication of hepatitis C virus in humans. Another important observation
reported in the above studies is the indirect effects of the miRNA miR-130a that inhibits the
expression of another miRNA miR-122. The miRNA miR-122 has been previously reported to
promote proliferation of hepatitis C virus (Jopling et al. 2005, 1577–1581).
Since SOCS3 is a competitive inhibitor of JAK kinases, miRNAs that inhibit SOCS3 would
have JAK–promoting effects. Conversely, those miRNAs that promote SOCS3 expression would
confer indirect inhibitory effects on JAK activation. Such an indirect role of miRNA miR-122
was recently reported to affect interferon treatment, which is the current standard of care to treat
hepatitis C viral infections. While overexpression of the miRNA miR-122 suppressed the
interferon-stimulated response element-mediated gene expression, silencing of miR-122 was
reported to enhanced interferon-induced functions, and as a consequence, the expression of
SOCS3 was reportedly affected (Yoshikawa et al. 2012, 637).
In another study, elevated levels of the miRNA miR-373 was reported in hepatitis C virus
infected human primary hepatocytes, and also in the liver biopsies obtained from human patients
(Mukherjee et al. 2015, 3356–3365). A direct target of miR-373 indicated in the above article is the
JAK1 transcript and the interferon regulating factor-9 and therefore, activation of STAT1 was found
to be decreased. Subsequently, knockdown of miR-373 in the hepatocytes was found to prevent the
expression of JAK1 and IRF9, and, therefore, the authors speculate that the hepatitis C virus induces
the expression of miR-373 in hepatocytes and this is required for its replication inside the cell.
42 JAK-STAT Signaling in Diseases

The Epstein–Barr virus infection of B lymphocytes has been recently shown to elevate the expression
levels of the miRNA miR-155, which contributes to immortalization (Lu et al. 2008, 10436–10443).
Subsequently, in HepG2 human hepatocellular carcinoma cell lines, it was reported that expression of
miR-155 upregulates the interferon inducible anti-viral response. However, overexpression of miR-155
was also found to inhibit SOCS1 expression and, as a consequence, STAT1 and STAT3 activation was
induced (Su et al. 2011, 354-422X-8-354). SOCS1 being a negative regulator of JAK-STAT pathway, the
above reports provide a clear understanding that the miRNA miR-155 enhances antiviral immune
response by activating JAK-STAT pathway, and this is required to prevent HBV infection in humans.
Diagnostic biomarkers, if detected early in the disease, can be very helpful in strategizing an effective
treatment plan to cure the disease. In this line, using the blood samples from 50 patients diagnosed with
HBV infections and hepatic fibrosis, circulating miRNAs were screened using miRNA microarray
(Zhang et al. 2015, 5647–5654). The authors reported that 12 miRNAs were differentially expressed, of
which 10 were overexpressed and 2 were downregulated. Among the target genes identified, 31 of them
were specific for JAK-STAT pathway. Although the study identified putative biomarkers for early
diagnosis of HBV-associated hepatic fibrosis, further validation studies are required.
As identified earlier in this manuscript, the role of miR-203 is evident in regulating different cellular
pathways in many disease conditions. Apart from the above-discussed studies, the role of miR-203 in
regulating viral infections was also reported. Within the 2500bp promoter region of miR-203, the
binding sites for two transcription factors, NFκB and ISGF3, were identified which are known to
regulate the JAK-STAT signaling pathway. ISGF3 is primarily involved in IFN-mediated regulation
of JAK-STAT pathway with STAT1, STAT2, and IRF9 as key mediators. Interestingly, in the IFN-
deficient cells that were infected with influenza A virus, increased expression of miR-203 was
observed. These reports suggest a direct correlation between the increased expression of miR-203
with influenza A virus infection; however, the same study also identifies indirect role of JAK-STAT
signaling during influenza A viral infection (Zhang et al. 2018, 6797-018-25073-9). IFNs are
commonly administered to treat viral infections. In both cancerous and non-cancerous liver cells,
treatment with IFN was reported to downregulate the levels of miR-1225-3p. Since viral infection
induces the production of endogenous IFN levels, Huh7 and 293T cells when infected with RNA
viruses, such as Hepatitis C Virus, Sendai Virus, and New castle Disease Virus, was also reported to
lower the levels of miR-1225-3p (Cheng et al. 2018, 5975–5986). Since IFN directly regulates JAK­
STAT pathway through its cell surface receptors, the above reports indicate IFN-mediated regulation
of JAK-STAT signaling by miR-1225-3p during viral infections.
The highly pathogenic influenza A virus, H5N1, showed more than 50% mortality rate in infected
people. In mice infected with H5N1, the expression of miR-324-5p was reported to be down-
regulated by Poly(IC), which is a viral pathogen-associated molecular pattern protein. By targeting
the RNA polymerase subunits PB1 and PB2, the miR-324-5p prevents H5N1 viral replication in the
host (Kumar et al. 2018). Further transcriptome analysis of the miR-324-5p identified one of its
direct target transcripts CUEDC2, which is a negative regulator of JAK-STAT signaling. It was also
shown that ectopic expression of miR-324-5p induces the expression of antiviral gene expression,
notably the IFNs. Taken together these reports show an essential role of different miRNAs in
regulating viral infections, primarily through IFN-mediated JAK-STAT signaling pathway compris­
ing different isoforms of JAK and STAT proteins.

3.7 Involvement of miRNAs in Other Diseases Regulated by


JAK-STAT Pathways
3.7.1 Spinal Cord Injury
The transcription factor SOX2 is critically required for tissue development and homeostasis in the
cells. In addition, for stem cells, SOX2 expression is required to maintain their self-renewability
and pluripotency. Just recently, the miRNA miR-103 was identified to prevent two important
JAK-STAT Pathway Regulation 43

cellular functions, autophagy and apoptosis, by regulating SOX2 expression. The miRNA miR-103
was shown to indirectly inhibit both MAPK-ERK and JAK-STAT pathways by upregulating SOX2
in PC12 cancer cells (Li et al. 2018, 292–300). Further in the same article, the authors showed that
in rats with spinal cord injury, delivery of the miRNA miR-103 agomir, increased the expression of
SOX2. The treatment with miR-103 agomir was also found to prevent the expression of Beclin-1,
Bax, activated caspase-3 and caspase-8, and in contrast, LC3-1, p62, and BcL-2 were reported to be
upregulated. Although whether miR-103 can directly regulate SOX2 expression remained unclear, it
was evident that SOX2 plays a critical role in tissue repair. Further, since CDK5R1 is a known
direct target of miR-103, it remains to be elucidated whether SOX2 is at a pivotal point between
CDK5R1 and JAK-STAT pathways, and their possible cross talk.
A mouse model of spinal cord injury due to contusion was reported as a feasible study to evaluate
axon growth and regeneration (Dai et al. 2018a, 582–589). In this mouse model, significantly reduced
expression of the miRNA miR-125b was noted due to spinal cord contusion. Both JAK1 and
STAT1 transcripts were identified as the direct targets of this miRNA miR-125b, suggesting
the direct role of JAK-STAT signaling in regulating spinal cord injury and axon regeneration.
The authors further extended the mice results to a rat model and reported that miR-125b reduces
protein levels of caspase3, which is a known activator of cellular apoptosis. The results presented
in the above study clearly indicate that miR-125b promotes axon growth and regeneration by
inhibiting the JAK-STAT signaling pathway. The same group also evaluated the role of another
miRNA miR-210 in the mice with spinal cord injury (Dai et al. 2018b, 6609–6615). A key
observation made in this study was the late recovery to grasping strength in mice with spinal cord
injury, compared to sham control group. The late recovery was found to correlate with the
reduced levels of the miRNA miR-210. Correlating with this observation, the mice group that was
administered with miR-210 mimic was found to show higher grip strength. In assessing the target
gene transcripts of miR-210, increased expression of JAK2 and STAT3 proteins was observed in
the control group. Their observations were further supported from the results seen in the mice
with miR-210 mimic group, which showed reduced JAK2 and STAT3 levels. The above two
studies point to the possibility of targeting JAK-STAT proteins to promote spinal cord injury. In
these lines, JAK or STAT inhibitors can be tested in animal models to show their effects in
promoting axon regeneration and recovery from spinal cord injury.

3.7.2 Eye Diseases


Eye transplant is vital to restoring vision in suitable patients. In efforts to address the inflammatory
responses in the retinal pigment epithelial cells (RPE) from donor eyes, the role of several cytokines
was recently evaluated (Kutty et al. 2010, 390–395). Three important cytokines, TNF-α, IL-1β, and
IFN-γ, were tested in RPE cells. The authors reported that all three were effective in inducing the
inflammatory responses in the RPE cells. While the effects of each cytokine were noticed in RPE cells
when treated individually, the cellular inflammation was reported to be exaggerated when RPE cells
were treated with all three cytokines at the same time. Importantly, the combined treatment was
shown to enhance the expression of miRNA miR-155, and the RPE cells when treated with the three
cytokines along with JAK1 inhibitor, the expression of miR-155 was drastically reduced. In addition,
the authors also showed that STAT1 activation was elevated in presence of the three cytokines and
the same was inhibited when the cells were treated with JAK1 inhibitor. These reports suggest that
miR-155 promotes JAK-STAT signaling to induce inflammation in the cells. However, additional
studies are required to precisely elucidate specific transcripts that are targeted by miR-155 and the
molecular intermediates that promote proinflammatory signaling in these cells.

3.7.3 Endothelial Cell Migration and Angiogenesis


Within this last decade, several studies have identified that targeting the programmed death ligand
1 “PD-L1” could be an efficient strategy to prevent progression of certain cancer types, especially
44 JAK-STAT Signaling in Diseases

with its initial success in treating melanoma patients (Wolchok et al. 2013a, 1–13, 2013b,
122–133). However, in many non-cancerous diseases, upregulation of cytokines is also a
characteristic feature such as in atherosclerosis where there is an increased expression of TNF­
α. Whether the increased expression of cytokines and growth factors in non-cancerous diseases
would have any impact on normal cells and tissues would enable strategizing a better treatment
plan. In these directions, using normal human lymphatic endothelial cells, the effects of the
cytokines TNF-α and IFN-γ were recently reported where the combined presence of TNF-α and
IFN-γ significantly elevated the expression of PD-L1 expression (Yee et al. 2017, 20683–20693).
A key finding in the above study was that TNF-α induced the expression of miRNA miR-155
and this effect was enhanced by treatment with IFN-γ. Further, the same effects on PD-L1 and
miR-155 expression were also reported in normal dermal fibroblasts. Importantly, it was
reported that the cytokines TNF-α and mostly the IFN-γ, initiates JAK-STAT pathways and
drives the activation of STAT1 and STAT3 contributing to the induced expression of miR-155
and PD-L1.
SOCS5 is another member of SOCS family, which is a negative regulator of JAK-STAT
pathway as it prevents activation of JAK. The miRNA miR-9 was reported to be increased in
different tumor types (Ma et al. 2010, 247–256; Shigehara et al. 2011, e23584). However, this
miRNA was found to target and inhibit SOCS5 expression by binding to its 3ʹUTR sequence
and relieves the inhibition on JAK activation (Zhuang et al. 2012, 3513–3523). Transfection of
the miRNA miR-9 into human and mouse endothelial cells was shown to induce phosphoryla­
tion of JAK1, JAK2, STAT1, and STAT3, and simultaneously inhibit SOCS5 expression.
Further, miR-9 was also shown to promote endothelial cell migration and angiogenic sprouting
in HUVECs.

3.7.4 Neuroprotective Functions


Besides the above-discussed biological effects of the miRNA, miR-155, its proinflammatory
functions through inhibition of SOCS1 were also reported in immortal mouse microglial cells
(Cardoso et al. 2012, 73–88). Upon transfection of the microglial cells with miRNA miR-155, the
expression of SOCS1 was significantly downregulated. On the contrary, knockdown of the
miRNA miR-155 not only upregulated the expression of SOCS1, but it also prevented induction
of nitric oxide synthase and inflammatory cytokines. It is to be noted that SOCS1 prevents JAK
and STAT activation. The authors report that when the microRNA miR-155 was inhibited
in microglial cells and treated with LPS, the LPS induced effects on the expression of TNF-α
and IL-6 were significantly reduced. These findings indicate that under chronic inflammatory
conditions, inhibition of miR-155 could confer neuroprotective effects and may prevent neuronal
cell death, which can be an effective treatment strategy for neurodegenerative diseases.
In the prion-infected mice brain, the levels of miRNA miR-146a was reported to be elevated.
Further, TLR2 and TLR4 activation also was found to hold strong correlation with the increased
expression of miR-146a (Saba et al. 2012, e30832). Interestingly, in the murine microglial cells, the
authors report that upon transfection with the miRNA miR-146a, the expression of the proin­
flammatory cytokine IL-6 was drastically reduced. However, the expression of the master
proinflammatory transcription factor, NFκB was affected in absence of miR-146a, in addition to
the JAK-STAT pathway mediators. Although these studies demonstrate the effect of miR-146a in
modulating the cellular inflammation, the cross talk between NFκB and JAK-STAT pathway is
still warranted.
Cerebral ischemia is a major cause of permanent damage to health that can lead to disability
and even death of the individual. Inflammation initiates several cellular mechanisms, which
eventually cause neuronal damage. In the cultured primary neuronal cells from mouse brain and
also in the mouse model for ischemic stroke induced due to middle cerebral artery occlusion,
targeting JAK2 was reported to confer neuroprotection against ischemic injury (Tian et al. 2018,
977–988). The cultured neuronal cells, when subjected to oxygen-glucose deprivation and
JAK-STAT Pathway Regulation 45

FIGURE 3.1 Regulation of JAK-STAT signaling pathway by microRNAs.

reoxygenation, appear to mimic the ischemic conditions. Expression of JAK2 and the downstream
proinflammatory molecules such as iNOS, MMP9, TNF-α, and IL-1β were reported in this
in vitro cell culture model. In this study, the miRNA miR-216a was identified as a direct target
for JAK2 and inhibition of JAK2 in the neuronal cells transfected with the miR-216a prevented
the expression of the above inflammatory mediators. Based on the above published reports,
targeting JAK2-mediated signaling appears to be a beneficial approach to protect neuronal
damage resulting from the ischemic injury.
Figure 3.1 shows known miRNAs that can directly target JAK and STAT transcripts and
prevent their translation. Also, specific miRNAs that can potentially promote JAK-STAT signal­
ing pathway by indirectly targeting the expression of proteins, such as the suppressors of cytokine
signaling proteins, which inhibit JAK1 and JAK2, were highlighted. In addition, the miRNAs,
which intercept the crosstalk between the JAK-STAT and PI3K–Akt pathways were shown in the
above figure.

3.8 Summary and Future Directions


The information presented here clearly shows that JAK-STAT pathway mediators promote patho­
logical progression of many human diseases. Several important miRNAs have been identified, which
have been proven to be key regulators of JAK-STAT signaling. In Figure 3.1, we have summarized
few miRNAs, which affect JAK-STAT pathway. Also, specific miRNAs were found to activate
JAK-STAT pathway in more than one disease conditions. Since inflammation and cell proliferation
are key cellular mechanisms that are associated with many diseases, and JAK-STAT pathway
promotes both these important cellular mechanisms, it appears very promising to identify and test
46 JAK-STAT Signaling in Diseases

those miRNAs that can modulate JAK-STAT pathway. As current bioinformatic tools were proven
to be effective in target identification and characterization of different miRNAs, high throughput
screening and cell-based assays would facilitate validation of the putative miRNAs to identify their
therapeutic benefits. Since miRNAs have gained entry into human clinical trials, their initial
evaluation through pre-clinical trials in animal models is already underway in many research
laboratories. It is anticipated that there would be a significant change in the therapeutic strategies
to treat different human diseases using miRNA-based medicine.

ACKNOWLEDGEMENTS
This work was supported by the State of Nebraska LB692 grant to CSB by Creighton University
and research grants R01 HL112597, R01 HL116042, and R01 HL120659 to DK Agrawal from
the National Institutes of Health, USA. The content of this book chapter is solely the responsi­
bility of the authors and does not necessarily represent the official views of the National Institutes
of Health or the State of Nebraska.

COMPETING INTERESTS
The authors declare no other relevant affiliations or financial involvement with any organization
or entity with financial interest or financial conflict with the subject matter or materials discussed
in this chapter. No writing assistance was utilized in the production of this manuscript.

REFERENCES
Boosani, C. S., K. Dhar, and D. K. Agrawal. 2015. “Down-Regulation of Hsa-miR-1264 Contributes to
DNMT1-Mediated Silencing of SOCS3.” Molecular Biology Reports 42 (9): 1365–1376.
Bott, A., N. Erdem, S. Lerrer, A. Hotz-Wagenblatt, C. Breunig, K. Abnaof, A. Worner et al. 2017.
“miRNA-1246 Induces Pro-Inflammatory Responses in Mesenchymal Stem/Stromal Cells by
Regulating PKA and PP2A.” Oncotarget 8 (27): 43897–43914.
Cardoso, A. L., J. R. Guedes, L. Pereira de Almeida, and M. C. Pedroso de Lima. 2012. “miR-155
Modulates Microglia-Mediated Immune Response by Down-Regulating SOCS-1 and Promoting
Cytokine and Nitric Oxide Production.” Immunology 135 (1): 73–88.
Cheng, M., Y. Niu, J. Fan, X. Chi, X. Liu, and W. Yang. 2018. “Interferon Down-Regulation of
miR-1225-3p as an Antiviral Mechanism through Modulating Grb2-Associated Binding Protein 3
Expression.” The Journal of Biological Chemistry 293 (16): 5975–5986.
Collins, A. S., C. E. McCoy, A. T. Lloyd, C. O’Farrelly, and N. J. Stevenson. 2013. “miR-19a: An Effective
Regulator of SOCS3 and Enhancer of JAK–STAT Signalling.” PLoS One 8 (7): e69090.
Dai, J., L. J. Xu, G. D. Han, H. L. Sun, G. T. Zhu, H. T. Jiang, G. Y. Yu, and X. M. Tang.
2018a. “MicroRNA-125b Promotes the Regeneration and Repair of Spinal Cord Injury through
Regulation of JAK/STAT Pathway.” European Review for Medical and Pharmacological Sciences
22 (3): 582–589.
Dai, J., G. Y. Yu, H. L. Sun, G. T. Zhu, G. D. Han, H. T. Jiang, and X. M. Tang. 2018b. “MicroRNA-210
Promotes Spinal Cord Injury Recovery by Inhibiting Inflammation Via the JAK–STAT Pathway.”
European Review for Medical and Pharmacological Sciences 22 (20): 6609–6615.
Dhar, K., K. Rakesh, D. Pankajakshan, and D. K. Agrawal. 2013. “SOCS3 Promotor Hypermethylation
and STAT3-NF-kappaB Interaction Downregulate SOCS3 Expression in Human Coronary Artery
Smooth Muscle Cells.” American Journal of Physiology.Heart and Circulatory Physiology 304 (6):
H776–H785.
Feng, H. J., W. Ouyang, J. H. Liu, Y. G. Sun, R. Hu, L. H. Huang, J. L. Xian, C. F. Jing, and M. J. Zhou.
2014. “Global microRNA Profiles and Signaling Pathways in the Development of Cardiac
Hypertrophy.” Brazilian Journal of Medical and Biological Research = Revista Brasileira De
Pesquisas Medicas E Biologicas 47 (5): 361–368.
JAK-STAT Pathway Regulation 47

Figueiredo Neto, M. and M. L. Figueiredo. 2017. “Combination of Interleukin-27 and MicroRNA for
Enhancing Expression of Anti-Inflammatory and Proosteogenic Genes.” Arthritis 2017: 6365857.
Gupta, G. K., K. Dhar, M. G. Del Core, W. J. Hunter 3rd, G. I. Hatzoudis, and D. K. Agrawal. 2011.
“Suppressor of Cytokine Signaling-3 and Intimal Hyperplasia in Porcine Coronary Arteries
Following Coronary Intervention.” Experimental and Molecular Pathology 91 (1): 346–352.
Jopling, C. L., M. Yi, A. M. Lancaster, S. M. Lemon, and P. Sarnow. 2005. “Modulation of Hepatitis
C Virus RNA Abundance by a Liver-Specific MicroRNA.” Science (New York, N.Y.) 309 (5740):
1577–1581.
Kumar, A., A. Kumar, H. Ingle, S. Kumar, R. Mishra, M. K. Verma, D. Biswas et al. 2018. “MicroRNA
Hsa-miR-324-5p Suppresses H5N1 Virus Replication by Targeting the Viral PB1 and Host
CUEDC2.” Journal of Virology 92 (19): doi: 10.1128/JVI.01057-18. Print 2018 Oct 1.
Kutty, R. K., C. N. Nagineni, W. Samuel, C. Vijayasarathy, J. J. Hooks, and T. M. Redmond. 2010.
“Inflammatory Cytokines Regulate microRNA-155 Expression in Human Retinal Pigment Epithe­
lial Cells by Activating JAK/STAT Pathway.” Biochemical and Biophysical Research Communica­
tions 402 (2): 390–395.
Li, G., T. Chen, Y. Zhu, X. Xiao, J. Bu, and Z. Huang. 2018. “MiR-103 Alleviates Autophagy and
Apoptosis by Regulating SOX2 in LPS-Injured PC12 Cells and SCI Rats.” Iranian Journal of Basic
Medical Sciences 21 (3): 292–300.
Li, H. W., Y. Xie, F. Li, G. C. Sun, Z. Chen, and H. S. Zeng. 2016. “Effect of miR-19a and miR-21 on the
JAK/STAT Signaling Pathway in the Peripheral Blood Mononuclear Cells of Patients with
Systemic Juvenile Idiopathic Arthritis.” Experimental and Therapeutic Medicine 11 (6): 2531–2536.
Li, S., X. Duan, Y. Li, B. Liu, I. McGilvray, and L. Chen. 2014. “MicroRNA-130a Inhibits HCV
Replication by Restoring the Innate Immune Response.” Journal of Viral Hepatitis 21 (2): 121–128.
Lu, F., A. Weidmer, C. G. Liu, S. Volinia, C. M. Croce, and P. M. Lieberman. 2008. “Epstein-Barr
Virus-Induced miR-155 Attenuates NF-kappaB Signaling and Stabilizes Latent Virus Persistence.”
Journal of Virology 82 (21): 10436–10443.
Lv, X., Y. Zhang, Y. Cui, Y. Ren, R. Li, and Q. Rong. 2015. “Inhibition of microRNA155 Relieves
Sepsisinduced Liver Injury through Inactivating the JAK/STAT Pathway.” Molecular Medicine
Reports 12 (4): 6013–6018.
Ma, L., J. Young, H. Prabhala, E. Pan, P. Mestdagh, D. Muth, J. Teruya-Feldstein et al. 2010. “miR-9, a
MYC/MYCN-Activated microRNA, Regulates E-Cadherin and Cancer Metastasis.” Nature Cell
Biology 12 (3): 247–256.
Martensen, P. M. and J. Justesen. 2004. “Small ISGs Coming Forward.” Journal of Interferon & Cytokine
Research: The Official Journal of the International Society for Interferon and Cytokine Research
24 (1): 1–19.
Mukherjee, A., A. M. Di Bisceglie, and R. B. Ray. 2015. “Hepatitis C Virus-Mediated Enhancement of
microRNA miR-373 Impairs the JAK/STAT Signaling Pathway.” Journal of Virology 89 (6): 3356–
3365.
Murayama, A., T. Date, K. Morikawa, D. Akazawa, M. Miyamoto, M. Kaga, K. Ishii et al. 2007. “The
NS3 Helicase and NS5B-to-3ʹX Regions are Important for Efficient Hepatitis C Virus Strain JFH-1
Replication in Huh7 Cells.” Journal of Virology 81 (15): 8030–8040.
Ranji, N., M. Sadeghizadeh, M. Karimipoor, M. A. Shokrgozar, R. Nakhaei Sistani, and S. H. Paylakhi.
2015. “MicroRNAs Signature in IL-2-Induced CD4+ T Cells and their Potential Targets.” Bio­
chemical Genetics 53 (7–8): 169–183.
Ren, W., S. Wu, Y. Wu, T. Liu, X. Zhao, and Y. Li. 2019. “MicroRNA-196a/-196b Regulate the
Progression of Hepatocellular Carcinoma through Modulating the JAK/STAT Pathway Via
Targeting SOCS2.” Cell Death & Disease 10 (5): 333.
Saba, R., S. Gushue, R. L. Huzarewich, K. Manguiat, S. Medina, C. Robertson, and S. A. Booth. 2012.
“MicroRNA 146a (miR-146a) is Over-Expressed during Prion Disease and Modulates the Innate
Immune Response and the Microglial Activation State.” PLoS One 7 (2): e30832.
Shigehara, K., S. Yokomuro, O. Ishibashi, Y. Mizuguchi, Y. Arima, Y. Kawahigashi, T. Kanda et al. 2011.
“Real-Time PCR-Based Analysis of the Human Bile microRNAome Identifies miR-9 as a Potential
Diagnostic Biomarker for Biliary Tract Cancer.” PLoS One 6 (8): e23584.
Su, C., Z. Hou, C. Zhang, Z. Tian, and J. Zhang. 2011. “Ectopic Expression of microRNA-155 Enhances
Innate Antiviral Immunity against HBV Infection in Human Hepatoma Cells.” Virology Journal 8: 354.
48 JAK-STAT Signaling in Diseases

Tian, Y. S., D. Zhong, Q. Q. Liu, X. L. Zhao, H. X. Sun, J. Jin, H. N. Wang, and G. Z. Li. 2018.
“Upregulation of miR-216a Exerts Neuroprotective Effects against Ischemic Injury through
Negatively Regulating JAK2/STAT3-Involved Apoptosis and Inflammatory Pathways.” Journal of
Neurosurgery 130 (3): 977–988.
Vimalraj, S. and N. Selvamurugan. 2014. “MicroRNAs Expression and their Regulatory Networks
during Mesenchymal Stem Cells Differentiation toward Osteoblasts.” International Journal of
Biological Macromolecules 66: 194–202.
Wolchok, J. D., F. S. Hodi, J. S. Weber, J. P. Allison, W. J. Urba, C. Robert, S. J. O’Day et al.
2013a. “Development of Ipilimumab: A Novel Immunotherapeutic Approach for the Treatment of
Advanced Melanoma.” Annals of the New York Academy of Sciences 1291: 1–13.
Wolchok, J. D., H. Kluger, M. K. Callahan, M. A. Postow, N. A. Rizvi, A. M. Lesokhin, N. H. Segal et al.
2013b. “Nivolumab Plus Ipilimumab in Advanced Melanoma.” The New England Journal of
Medicine 369 (2): 122–133.
Wu, D. P., J. L. Zhang, J. Y. Wang, M. X. Cui, J. L. Jia, X. H. Liu, and Q. D. Liang. 2017. “MiR-1246
Promotes LPS-Induced Inflammatory Injury in Chondrogenic Cells ATDC5 by Targeting
HNF4gamma.” Cellular Physiology and Biochemistry: International Journal of Experimental Cel­
lular Physiology, Biochemistry, and Pharmacology 43 (5): 2010–2021.
Yao, R., Y. L. Ma, W. Liang, H. H. Li, Z. J. Ma, X. Yu, and Y. H. Liao. 2012. “MicroRNA-155
Modulates Treg and Th17 Cells Differentiation and Th17 Cell Function by Targeting SOCS1.”
PLoS One 7 (10): e46082.
Yee, D., K. M. Shah, M. C. Coles, T. V. Sharp, and D. Lagos. 2017. “MicroRNA-155 Induction Via
TNF-Alpha and IFN-Gamma Suppresses Expression of Programmed Death Ligand-1 (PD-L1) in
Human Primary Cells.” The Journal of Biological Chemistry 292 (50): 20683–20693.
Yoshikawa, T., A. Takata, M. Otsuka, T. Kishikawa, K. Kojima, H. Yoshida, and K. Koike. 2012.
“Silencing of microRNA-122 Enhances Interferon-Alpha Signaling in the Liver through Regulating
SOCS3 Promoter Methylation.” Scientific Reports 2: 637.
Zafra, M. P., C. Mazzeo, C. Gamez, A. Rodriguez Marco, A. de Zulueta, V. Sanz, I. Bilbao, J. Ruiz-
Cabello, J. M. Zubeldia, and V. Del Pozo. 2014. “Gene Silencing of SOCS3 by siRNA Intranasal
Delivery Inhibits Asthma Phenotype in Mice.” PLoS One 9 (3): e91996.
Zhang, C. S., Y. Lin, F. B. Sun, J. Gao, B. Han, and S. J. Li. 2019. “miR-409 Down-Regulates JAK–Stat
Pathway to Inhibit Progression of Liver Cancer.” European Review for Medical and Pharmacologi­
cal Sciences 23 (1): 146–154.
Zhang, Q., M. Xu, Y. Qu, Z. Li, Q. Zhang, X. Cai, and L. Lu. 2015. “Analysis of the Differential
Expression of Circulating microRNAs during the Progression of Hepatic Fibrosis in Patients with
Chronic Hepatitis B Virus Infection.” Molecular Medicine Reports 12 (4): 5647–5654.
Zhang, S., J. Li, J. Li, Y. Yang, X. Kang, Y. Li, X. Wu, Q. Zhu, Y. Zhou, and Y. Hu. 2018. “Up-Regulation
of MicroRNA-203 in Influenza A Virus Infection Inhibits Viral Replication by Targeting DR1.”
Scientific Reports 8 (1): 6797.
Zhao, C., Y. Wang, H. Jin, and T. Yu. 2017. “Knockdown of microRNA-203 Alleviates LPS-Induced
Injury by Targeting MCL-1 in C28/I2 Chondrocytes.” Experimental Cell Research 359 (1): 171–178.
Zhu, J., K. Yao, J. Guo, H. Shi, L. Ma, Q. Wang, H. Liu et al. 2017. “miR-181a and miR-150 Regulate
Dendritic Cell Immune Inflammatory Responses and Cardiomyocyte Apoptosis Via Targeting
JAK1-STAT1/C-Fos Pathway.” Journal of Cellular and Molecular Medicine 21 (11): 2884–2895.
Zhuang, G., X. Wu, Z. Jiang, I. Kasman, J. Yao, Y. Guan, J. Oeh et al. 2012. “Tumour-Secreted miR-9
Promotes Endothelial Cell Migration and Angiogenesis by Activating the JAK–STAT Pathway.”
The EMBO Journal 31 (17): 3513–3523.
4
JAK-STAT Signaling in Asthma and Allergic
Airway Inflammation

Amina Abdul Qayum, Tristan Hayes, and Mark H. Kaplan


Department of Pediatrics and Herman B Wells Center for Pediatric Research
Indiana University School of Medicine
Indianapolis, Indiana

4.1 Introduction
Classical asthma is an inflammatory disease mediated by TH2 and TH9 cells characterized by
increased serum IgE, eosinophilia, and responsiveness to steroids. Non-canonical asthma, which
affects less than 10% of all asthmatics, is thought to be promoted by TH17 cells and is non-atopic,
characterized by steroid-resistant neutrophilia.1 Cytokines that are well known to implement the
asthmatic phenotype include IL-4, IL-5, IL-9, IL-13, IL-17, TSLP, and IL.-312,3 These cytokines
signal through the JAK-STAT (see Table 4.1) pathway to mediate communication between cells and
their environment. In this chapter, we will discuss the current understanding of JAK-STAT signaling
in multiple cell types as it relates to asthma and allergic lung inflammation in animal models.

4.2 Janus Kinase


Janus kinase (JAK) proteins are a family of intracellular non-receptor tyrosine protein kinases
(PTKs) that aid in signal transduction of JAK-STAT pathways. There are four known JAK kinases:
JAK1, JAK2, JAK3, and TYK2. The structures of these kinases contain seven homology domains
(JH1-JH7). The kinase domain (JH1) has catalytic activity while the pseudo-kinase domain (JH2)
located at the carboxy terminus end, does not have catalytic activity. JH2 domains interact with and
regulate the JH1 domain.4,5 The kinase and pseudo-kinase domains are the defining feature of the
JAK family of proteins. This name is derived from the Roman God of duality, Janus, who is depicted
as having two heads; one looking to the past and one looking to the future. The JH3–JH4 domains
contain homology to SH2 domains and putatively play a structural role and stabilize JAK protein
conformation. The JH5–JH7 domains are located at the amino-terminus and serve as the FERM
proteins that are critical for receptor binding of JAKs.6,7
When a ligand binds to its receptor, JAKs associated with that receptor phosphorylate each
other as well as the cytokine receptor. The phosphorylated receptor serves as a docking site for
STAT proteins. When the STAT proteins bind to the receptor, they become phosphorylated by
JAKS. Activated STATs can then translocate to the nucleus where they directly bind DNA and
promote RNA transcriptoin. There is also evidence for the role of nuclear JAK proteins in cell
signaling. Some studies have shown that JAKs can modify other non-STAT transcription factors
and modify histone proteins.8,9 However, the main function defined for JAKs is in the activation
of STAT proteins.

49
50 JAK-STAT Signaling in Diseases

TABLE 4.1
Th2 Cytokine Usage of JAK-STAT Proteins
JAK/STAT Role in Signaling Mutation Phenotype References

JAK1 IL-4, IL-13, IL-5, IL-9, KO prenatally lethal 6


TSLP, IL-31
JAK2 IL-5, TSLP, IL-31 KO embryonically lethal 6,7
JAK3 IL-4, IL-13, IL-9 KO causes SCID (similar in humans) 12,8,11
Defective B and T cell maturation
Tyk2 (IFNs, IL-12) KO susceptible to infections 10
Defective STAT3 activation
STAT1 (IFNs), IL-5, IL-31, IL-9 Defective in Type I/II IFN signaling 18,21,19
STAT2 IL-5 KO phenotype not yet reported 18,21,19
STAT3 IL-5, IL-31, IL-9 KO embryonically lethal and mutation causes HIES 27
STAT4 (IL-12) Defective Th1 Development 33,34,35
Enhanced Th2 development
STAT5a/b IL-5, TSLP, IL-31, IL-9 STAT5a/b: defects in hematopoietic cell development and 33,34,35,38,40
immune cell signaling
STAT6 IL-4, IL-13 Defective Th2 and Th9 development 46
Protected from AHR 132, 178
KO, knock out; AHR, airway hyperresponsiveness; TSLP, thymic stromal lymphopoietin; SCID, severe combined
immunodeficiency; IFN, interferon. Brackets indicate canonical activators of the pathway that are not Th2 cytokines.

4.2.1 JAKs and Asthma


The importance of JAK proteins in human diseases is evident by the presentation of autoimmune
and inflammatory disorders in humans with JAK mutations. Somatic missense mutations in JAKs
are linked to the development of malignancies. Genome-wide association studies (GWAS) have
not yet identified small nucleotide polymorphisms (SNPs) in JAK genes linked to asthma.8
In mice, the loss of JAK1 or JAK2 is lethal, respectively, at perinatal or embryonic stages of
development. Based on this evidence, loss of function mutations in JAK1 and JAK2 are predicted
to have high level of intolerance in humans.
Tyk2−/− animals show increased susceptibility to viral and intracellular infections.10 The deficiency
in Tyk2 affects Type II interferon activity and IL-12 signaling.11 Only a few patients have been
identified with Tyk2 mutations and all of them show high susceptibility to infections.11 One patient
with a homozygous recessive Tyk2 mutation was diagnosed with hyper-IgE syndrome (HIES), which
is an immune deficiency characterized by high levels of IgE, eczema, and recurrent lung infections.11
HIES patient cells are also defective in several cytokine signaling pathways (IL-6 and IL-12).10,11 Of
note is that Tyk2−/− mice do not develop HIES despite defects in cytokine signaling (IL-12).10 This
case serves as a good example of differences in JAK signaling between human and animal models,
though it could also indicate differences between loss of the entire gene and hypomorphic mutations
in patients, highlighting the complexity behind developing and testing treatments.
JAK3 is unique in that it is not expressed in all cell types like the other three JAKs. JAK3
expression is predominantly present in hematopoietic cell lineages and it only associates with IL-2
common gamma chain receptor. JAK3−/− animals have defective T and B cell maturation.
Mutations in JAK3 cause autosomal recessive severe combined immunodeficiency syndrome
(SCID). Human mutations of JAK3 are limited to defects in the immune system, which has
made targeted inhibitors of JAK3 promising immunosuppressants.8,9,12

4.2.2 JAKinibs in Asthma


The criticality of JAKs in cytokine signaling has led to the development of numerous JAK
inhibitors; both specific to certain JAKs and broad tyrosine kinase inhibitors. Some of these
Asthma and Allergic Airway Inflammation 51

TABLE 4.2
Use of JAK-STAT Protein Inhibitors in Mouse Models of Asthma
Inhibitor Targeted Protein Mouse Asthma Model References

TyrA1 All JAKs ↓STAT3 activation 15


↓Eosinophilia
WHI-P97 JAK3 ↓Eosinophilia 14
↓AHR
VR588 All JAKs ↓STAT3 activation 16
↓AHR
↓Immune cell infiltrate into lung
Decoy ODN STAT1 ↓BAL infiltrate, AHR, IL-5 56
C1889 STAT3 ↓Th2, Th17 cell development 57
R256 JAK1, JAK3 ↓Eosinophilia 62
↓AHR
↓Mucus production
AS1517499 STAT6 ↓RhoA Expression 196
Leflunomide STAT6 ↓RhoA Expression 197
AHR, airway hyperresponsiveness; BAL, bronchoalveolar lavage.

inhibitors are FDA approved and are being prescribed for conditions like rheumatoid arthritis
(Ruxolitinib, Tofacitinib, Oclacitinib), where they show considerable efficacy.13 These inhibitors
are in clinical trials for their potential in the treatment of other immune-related diseases as well.13
While JAK inhibitors have not yet been tested in patients with asthma, mouse model studies of
asthma have employed JAK inhibitors (see Table 4.2). One study shows that inhibiting JAK3
specifically (WHI-P97) reduces leukotriene synthesis in mast cells and in vivo administration can
prevent eosinophilia and airway hyperresponsiveness (AHR).14
A pan-JAK inhibitor (TyrA1) has been used which resulted in reduced STAT3 activation and
eosinophilia during a mouse model of allergic asthma.15 Another pan JAK inhibitor (VR588) also
resulted in reduced STAT3 activation, AHR, and immune cell infiltration in mice during allergic
asthma.16 These studies provide clear evidence that JAKs are important in orchestrating asthma
and can be targeted for inhibition. However, it is still not clear how they will affect asthmatic
patients.

4.3 STATs and Asthma


STATs are latent cytoplasmic proteins classified as transcription factors. There are seven known
mammalian STAT proteins (STAT1, STAT2, STAT3, STAT4, STAT5a, STAT5b, STAT6) (see
Table 4.1). The structure of STAT proteins allows them to serve a dual role of both signal
transduction and DNA binding. Sequentially from the amino terminus end to the carboxy
terminus, these proteins have a STAT dimerization domain, coiled-coil domain, DNA binding
domain, SH2 domain, and transactivation domain. STAT proteins use their SH2 domains to bind
phosphorylated cytokine receptors. STATs are then phosphorylated by JAK kinases and released
from the receptor. This allows STATs to form homo or hetero dimers, which are poised for
translocation into the nucleus and transcriptional activation.17

4.3.1 STAT1 and STAT2


The overall importance of STAT1 and STAT2 for asthma is unclear. Type I, Type II, and Type III
interferons signal through STAT1 and STAT2. While some evidence exists that interferons may
52 JAK-STAT Signaling in Diseases

perpetuate asthma in general, Interferons are thought to be protective in asthma development


because they skew the immune system to Th1-like responses.18–20
Through the use of knockout animals and administration of recombinant IFNs during allergic
asthma, studies have shown that IFN-λ can reduce IL-5-, IL-13-, and IL-17-positive T cells and
dampen allergic asthma in mice.19–21 One study found that indirect activation of STAT1 (through
interferons or IL-12) can be protective in allergic asthma.22 There is also evidence that STAT1
induces pulmonary chemokines indicating that under some conditions it may not play a protective
role in mouse models of asthma.23 Evidence for the upregulation of STAT1 activity in epithelial
cells of asthmatic patients has also been documented.24

4.3.2 STAT3
Loss of function mutations in STAT3 in humans is the underlying cause of dominant negative
hyper-IgE syndrome (HIES), while gain of function mutations in STAT3 is associated with the
development of autoimmunity and immunodeficiencies. No association has been found between
SNPs or mutations in STAT3 and occurrences of asthma.25 It has been demonstrated that STAT3
is a positive regulator of migration and function of eosinophils from asthmatic patients.26
Germ-line Stat3−/− animals are difficult to assess because they die early in embryonic develop­
ment. However, conditional mutant mice that lack STAT3 in specific tissues have been used to
address the role of STAT3 in allergic inflammation. The importance of STAT3 in house dust mite
(HDM) models of asthma has been demonstrated using STAT3 airway epithelial cell conditional
mutant animals. These animals have reduced airway eosinophilia, lung TH2, and TH2 cytokine
levels in a chronic asthma model.15
STAT3 is known to play an important role in the development of TH2, TH9, and TH17 cells.
Mice with Stat3−/− T cells fail to develop allergic asthma. STAT3 can bind to TH2 cytokine loci
and this process is critical for STAT6 induced TH2 cell development.27 STAT3 is also critical for
inhibiting stability of TH9 cell development.28 In TH17 cells, STAT3 can activate several cytokines
and this process is well documented. T cells that are Stat3−/− have reduced RORγt,
a transcription factor that is critical for TH17 cell development. Similarly, TH17 cells fail to
develop in HIES patients.27,29 However, in contrast to the mouse system, STAT3 is required for
human TH9 development.29

4.3.3 STAT4
STAT4 is expressed in lymphoid and myeloid cells.30 It is implicated in the development and
regulation of TH1, TH2, and follicular helper T cells.31–35 STAT4 is capable of exerting cytokine
regulatory effects in other cell types such as regulatory T cells, TH17 cells, mast cells, and natural
killer cells.36–39 Beyond its role in cell development, STAT4 is associated with increasing the
severity of airway diseases including chronic obstructive pulmonary disorder (COPD), allergic
rhinitis, and asthma.40–43 At present, however, the direct role that STAT4 plays on the develop­
ment of these diseases remains under-investigated. The role of STAT4 on the regulation of several
cytokines important for the development of airway diseases such as asthma, has been explored
using mouse models of asthma and airway hypersensitivity.
STAT4 is a transcription factor that was initially described as being a master controller of the
development of TH1 cells. STAT4, which can be activated by IL-12, is capable of inducing effects
ranging from increased IFN-γ production, lymphocyte proliferation, and enhanced natural killer
cell cytotoxicity.33,35 During viral infections, STAT4 binds to the IFN-γ gene suggesting that
STAT4 alone is capable of regulating the IFN-γ pathway.44
As asthma is an airway disease with a complex interplay between TH1, TH2, and TH17 cells, the
cytokine milieu can have conflicting effects on TH cell development. IL-12 promotes the differ­
entiation of naïve T cells into TH1 cells while suppressing the generation of TH2 cytokines such as
IL-4 and IL-5. However, TH2 cytokines such as IL-4 inhibit the development of TH1 cells. Much
Asthma and Allergic Airway Inflammation 53

of this interplay can be attributed to transcriptionally activated T-bet which induces IL-12
receptor beta 2 subunit (IL-12Rβ2) expression and subsequently IL-12-dependent STAT4 activa­
tion. T-bet conversely represses the expression of several transcription factors including GATA-3.
In an environment with IL-4, IL-5, and IL-13, IFN-γ production and IL-12Rβ2 gene expression
are suppressed.45–48 TH17 cells are additional players in some asthma models and IL-12-induced
STAT4 activation and the induction of T-bet can drive them to an IFN-γ producing phenotype
further suppressing TH17 cell development.49–54

4.3.3.1 The IL-12/STAT4/IFN-γ Pathway in Mouse Models of Asthma


In mouse models of allergen-induced asthma, direct treatment with IL-12 can inhibit the
development of asthma-like symptoms, including airway hyperresponsiveness and eosinophilic
infiltration of the airways.55–57 Several studies elucidate the mechanism that drives the control
of STAT4 and IL-12 on airway hyperresponsiveness in allergen-induced asthma.58 Mice
deficient in STAT4, on both the C57BL/6 and BALB/C background, develop airway hyperre­
sponsiveness and airway eosinophilia.59,60 When these mice were given an IL-12 blocking
antibody during allergen re-challenge, airway hyperesponsiveness was reduced.61 In this model,
Stat4−/− mice showed decreased numbers of peribronchial eosinophils as compared to
controls. As expected, when profiling the airways for cytokines, IL-12 was lower as compared
to WT controls, perhaps suggesting that the lack of IL-12-induced STAT4 directly affects
eosinophils.61 IFN-γ was not altered in this model, suggesting an IFN-γ-STAT4-independent
process in some models of allergen-induced asthma. One such mechanism has been
described.61–63 Interestingly, in other studies using WT mice, when IL-12 is applied directly to
the airways of OVA-sensitized mice, IFN-γ levels are increased after antigen challenge. In this
model, eosinophil recruitment is still inhibited.58 Additionally, when mice are treated with TH2
cytokine reducing compounds, such as CpG Oligodeoxynucleotide, the ratio of IFN-γ and
IL-12 increases as compared to IL-5. This, in turn, delays the development of asthma-like
symptoms.64 Thus in these studies, the role that IL-12/STAT4 plays in asthma could be
beneficial.
In other mouse models of allergen-induced asthma, STAT4 contributes to airway hyperrespon­
siveness. However, its contribution compared to IFN-γ appears dependent upon concentration of
allergen used. For example, in one study, low dosage of LPS (0.1 µg) induced type 2 asthma
(airway hyperresponsiveness and eosinophilic infiltration) while high dosages (10 µg) induced type
1 asthma (airway hyperresponsiveness and non-eosinophilic infiltration). In the high dose model,
the asthma was not able to develop when using IFN-γ-deficient mice. Levels of lung inflammation,
IgG2a production, and airway infiltration by IFN-producing cells (CD8+ helper T cells) were
decreased. Interestingly, when using Stat4−/− mice, high dosages of LPS did not induce pulmon­
ary inflammation.65 A separate study utilizing cockroach antigen-treated control and Stat4−/−
mice showed that pulmonary levels of chemokines CCL5, CCL6, CCL11, and CCL17 were
reduced when compared to WT mice. Additionally, and unlike the previous model, TH2 cytokines
including IL-4 and IL-13 were not reduced.61 Allergen-induced airway hyperresponsiveness was
found to be impaired by STAT4-deficiency in a methacholine-challenge airway hyperresponsive­
ness model. Though IL-12 levels remained the same between WT and Stat4−/− groups, lung
infiltration by IFN-γ producing cells was only enhanced in WT mice.66 Thus, most data suggest
that IL-12-signaling and STAT4-dependent IFN-γ production are critical to controlling airway
hyperresponsiveness, but in some models they can be dispensable or even required for develop­
ment of disease.

4.3.3.2 STAT4 in Patient Asthma


Several population-based studies have found that IL-12 levels are reduced in asthmatics and that
polymorphisms in the IL-12 gene can serve as a risk factor for the development of asthma.67,68
54 JAK-STAT Signaling in Diseases

Following inhaled corticosteroid treatments, IL-12 levels are increased in these patients.69 How­
ever, when using high dose-inhaled corticosteroids on populations with defects in the STAT4 gene,
asthma symptoms do not improve.43
Additionally, human airway disease studies have mixed conclusions in correlating IFN-γ and
disease severity.70–76 For example, sputum IFN-γ expression is higher in severe asthmatics as
compared to mild to moderate asthmatics.65 In a study examining discordant twins, hypermeth­
lyation of IFN-γ was associated with asthma severity.77 Another study found hypermethlyation of
the IFN-γ promoter to be associated with the increased risk of the development of occupational-
related asthma.78 These suggestions indicate that the role of IFN-γ may be more subtle and
context-dependent.
In the human population, polymorphisms of STATs, including STAT4 and STAT6, are
associated with airway diseases including asthma and COPD.79,80 In studies using a Japanese
population of asthmatics, patients homozygous for the STAT4 SNP (rs925847) display reduced
lung function and increased bronchial airway remodeling.43 In a study examining an asthmatic
Korean population, that same SNP was associated with an increased risk of the development of
allergic asthma.81 In studies of children, STAT4 SNPs (rs16833215 and rs4853546) were asso­
ciated with childhood wheezing and increased asthmatic exacerbations.82 Finally, in studies
incorporating IFN-γ, STAT4, and STAT6 polymorphisms, SNPs in any of these genes were
correlated with the development of asthma.83 Taken together, these studies suggest that defects
in the STAT4-IFN-γ pathway could lead to the development of asthma and increased severity of
symptoms.

4.3.4 STAT5
There are two genes encoding STAT5 proteins: Stat5a and Stat5b which are 96% similar in
sequence but have some distinct biological functions. Stat5a/b−/− animals have prenatal
lethality and severe immunodeficiency similar to JAK3−/−.84 In humans, STAT5b deficiency
causes autoimmune disease in part due to dysfunctional T regulatory cells.85 Patients with severe
refractory asthma (SRA) were found to have lower levels of STAT5a in their peripheral blood
mononuclear cells.86
STAT5 proteins can be activated via JAK and Src kinases. It has been well documented that
STAT5 plays an important role in the development of cells that drive airway disease such as TH2,
TH9, ILC2s, mast cells, and eosinophils.87–89 STAT5 is critical in mast cell development and
IgE-mediated mast cell function.87 In TH9 cells, IL-9 gene transcription is in part dependent on
STAT588,90,91 STAT5 activation, through IL-2 signaling, has been shown to induce T cell
proliferation in asthma models.92

4.3.4.1 STAT5, IL-5, and IL-2


Eosinophilic lung inflammation and high peripheral blood eosinophil numbers have been asso­
ciated with severity of asthma in humans. Eosinophils are also implicated in airway remodeling
because they are the source of remodeling molecules such as tumor growth factors (TGF), VEGF,
and metalloproteinases.89 IL-5 is undoubtedly an important mediator of asthma because it is
important for the maturation and migration of eosinophils. In the asthmatic lung milieu, IL-5 is
primarily produced by T cells, eosinophils, mast cells, and innate lymphoid cells. In humans, high
levels of IL-5 protein can be found in the mucosa of asthmatic airways.93 Studies of anti-IL-5
treatment (Mepolizumab) have shown reduced disease exacerbation in severe asthmatics with
eosinophilia. Anti-IL-5 treatment in these patients reduced peripheral blood and sputum eosino­
phil counts.94–96
Ovalbumin-challenged Il5−/− mice show decreased numbers of circulating eosinophils and
airway eosinophilia.97,98 Conversely, inhalation of IL-5 has been shown to increase eosinophils
and airway hyperresponsiveness. IL-5 primarily signals through the JAK2/STAT5 pathway. It has
Asthma and Allergic Airway Inflammation 55

been demonstrated that Stat5a−/− and Stat5b−/− animals have diminished antigen-induced
eosinophil recruitment in the airways. Interestingly, this effect is due to both diminished IL-5
responsiveness of eosinophils and defects in antigen-specific TH2 cell proliferation.99,100 TH2 cell
differentiation is thought to be orchestrated by IL-2 signaling, which primarily signals through
STAT5. It has been shown that constitutively active STAT5a, independent of STAT6, can increase
accessibility of the IL-4 gene leading to TH2 cell differentiation.101 Consistent with this concept,
mice that are Stat5a−/− Stat6−/− have severely decreased allergic airway inflammation with low
eosinophilia and TH2 cell differentiation compared to Stat6−/− alone.102

4.3.4.2 STAT5 and TSLP


TSLP is a type I cytokine that signals through JAK2/JAK1 to activate STAT5 (and other STATs
to a lesser degree).103 In the lung, TSLP is primarily expressed in epithelia and stromal cells and
some nonstructural cells such as DCs.104 TSLP supports maturation of B and T cells in addition
to modulating many immune cells that play a role in development of lung inflammation like DCs,
mast cells, and eosinophils. TSLPR-deficient animals have been used to demonstrate that without
TSLP signaling, mice do not develop allergic asthma. Inflammation can be restored in these
animals when wild type T cells are reconstituted, emphasizing the importance of TSLP signaling
for T cells.105,106 It has also been demonstrated that TSLP inhibits the development of tolerance
to an allergen in the lung. TSLP can suppress development of antigen-specific human and mouse
T regulatory cells.107,108

4.3.5 STAT6
In mice and humans, IL-4 induces the tyrosine phosphorylation and DNA binding activity of
STAT6109,110 IL-13 activates STAT6 in cells that express the receptor including macrophages.46,111,112
Some evidence suggests that IL-15 is additionally capable of activating STAT6; however, more
investigations are warranted.113–115 Activated STAT6 has roles in the regulation of several genes
including the TH2 transcription factor GATA3, IL-4Rα receptor, eotaxin-3, CCL17, and
FIZZ.146,115–124 STAT6 also inhibits the expression of CD40 and E-selectins.125,126
IL-4 and IL-13 are cytokines that induce the activation of STAT6. In asthma, these cytokines
play numerous roles. For example, activated TH2, NK, and mast cells secrete IL-4 and IL-13 to
drive the activation of B-cells. Activated B-cells initiate an IgE-producing phenotype, which allows
mast cell priming.127–130 In asthmatics, both cytokines are increased in airway smooth muscle
cells.131 These cytokines promote goblet cell metaplasia which enhances mucus secretion and airway
hyperresponsiveness.127,130,132,133 Therefore, it is of interest to identify the link between IL-4 and IL-13
with STAT6 in asthma. Indeed, asthma patients treated with the IL-4/IL-13 signaling inhibitor,
dupilimab, experience decreased asthma exacerbations and improved lung functions.134
STAT6 plays several roles in helper T-cell development. While it mainly skews the differentia­
tion of naïve T-cells towards a TH2 producing phenotype, it can regulate both TH9 and TH17 cell
development.35,46,135–137 In airway diseases, STAT6 affects numerous cell types. For example, in an
Alternaria-allergen challenge model of asthma, STAT6 controlled eosinophilia.138 Another study
linked the control of eosinophilia by STAT6 to PARP1, a poly (ADP-ribose) polymerase.139
In asthmatics, STAT6 positively correlates with increased serum IgE levels.140 Furthermore,
several studies indicate that increased STAT6 levels also correlate with more severe forms of
allergic rhinitis and asthma.141–145 In patients who undergo glucocorticosteroid treatments,
STAT6 controls cytokine secretion in cell types such as innate lymphoid cells (ILC2s).146

4.3.5.1 STAT6 and IL-4 in Patient Asthma


Observations of induced sputum in asthmatics reveal that STAT6 and the IL-4 receptor are
expressed more highly in these patients than in control subjects.147 Mutations in the IL-4 receptor
56 JAK-STAT Signaling in Diseases

are wide ranging and can result in increased serum IgE. Several mutations are linked with the
increased incidence of atopic asthma and defective STAT6 DNA binding (Q576R and
S503P).148–153 For example, a mutation in an extracellular region of IL-4R (V50) leads to prolonged
STAT6 phosphorylation.153 Another IL-4R mutation (Ala57Thr) is associated with decreased atopy
among patients in Greenland.150
Several STAT6 variants are associated with asthma risk.154 The variant rs324015 has a
protective effect on atopic asthma. Variant rs7180246 is associated with an increased risk of
asthma in all but an Indian population.143,154 In Caucasians, the variant rs324011 is associated
with asthma risk.154 Variant rs324011 though not associated with asthma risk, is associated with
recurrent wheezing in early childhood.155 A recently characterized STAT6 variant, rs167769, is
associated with asthma risk in a multi-ancestry study.156

4.3.5.2 Pharmaceutical and Biological Inhibition of STAT6


Researchers have used various inhibitors to block the IL-4/STAT6 pathway. Fasudil, an
approved clinical drug that can block genes such as RhoA/Rho kinase from enhancing
hypermucus secretion in asthmatics, was used in mouse studies to further clarify the role of
STAT6 in asthma. In an OVA-challenge mouse model, secreted IL-4 was decreased in Fasudil­
treated animals compared to untreated wild types. Importantly, this correlated with a decrease
in both phosphorylated STAT6 and STAT6 protein levels as compared to OVA-challenged
untreated controls.157
Another drug, heparin, limits the amount of secreted cytokines from asthmatic cells and prevents the
development of asthma in mice, sheep, and pigs.158–161 In humans, it limits bronchoconstriction.162 In
an OVA-sensitized mouse model of asthma using sulfated non-anticoagulant low molecular weight
heparin (S-NACH), a low molecular weight form of heparin, S-NACH acted as an inhibitor of STAT6.
This led to reduced secreted IL-4. Transcription factor protein amounts, such as GATA3, were also
reduced.163
Reductions in protein expression levels of IL-4 in mouse lungs occur when STAT6 siRNA is
used in asthmatic models.164 Similarly, in models of allergic rhinitis, sneezing and nasal rubbing in
STAT6 siRNA-treated mice are decreased as compared to untreated controls. When submandib­
ular lymph node cells were stimulated with OVA in vitro, secreted IL-4 amounts were undetected
in cells originating from the STAT6 siRNA-treated mice.165 These results corroborate with those
from similar studies. In chronic fungal asthma mouse models using Stat6−/− mice, IL-4 levels
increased much slower than in WT mice.166 The lack of complete abolishment of IL-4 suggests
that STAT6 does not prevent all features of asthma from developing.167 Chemokines such as
MCP-1, RANTES, and eotaxin were altered compared to WT animals. This confirmed the results
of several Stat6−/− in vitro studies.166,168,169 Studies using STAT6-deficient mice and acute and
chronic challenge with OVA, also show decreased amounts of IL-4 in BAL.167
Several chemical compounds provide insights into the role of STAT6 and IL-4 in allergic airway
disease. In rat allergic airway models using YM-341619 hydrochloride, a compound that exclu­
sively suppresses the differentiation of T cells to TH2 cells, STAT6 expression was suppressed.
When the compound was used on cultured T-cells, IL-4 production was decreased. Airway
hyperresponsiveness was similarly decreased when the compound is administered orally in a DNP-
Ascaris asthmatic-mouse model.170 This finding mimics the role of Prednisolone, an orally
available corticosteroid, in altering IL-4 levels in asthma. In an OVA sensitization and methacho­
line-challenge model using BALB/c mice, the use of Boswellic acid led to decreased IL-4 secretion,
phopho-STAT6, and GATA3. Overall, airway inflammation was attenuated.171 In a study using
tetrahydrocurcumin (THC), a major metabolite of Curcumin, a pigment from the Indian spice
turmeric, OVA-induced asthmatic mice displayed decreased asthmatic symptoms, such as nasal
rubbing, tissue eosinophilia, macrophage infiltration, and mucus production. Cytokine levels in
BAL fluid such as IL-4, IL-5, and IL-13 were reduced compared to OVA-challenge THC-
untreated mice. In examining protein secretion and phosphorylation levels, the data suggested
Asthma and Allergic Airway Inflammation 57

that these decreases were due to decreased expression of IL4 receptor, STAT6, and GATA3,
further suggesting a role for these proteins in asthma.172
Natural anti-inflammatory compounds such as ursolic acid have been used in lung inflamma­
tion treatments. In one study, ursolic acid prevented the development of airway eosinophilia in
OVA-challenged mice. Levels of several cytokines such as IL-13 and IL-17 were also decreased as
compared to wild type controls. Interestingly, when examining the levels of transcription factor
expression, both STAT6 and GATA3 were decreased compared to controls.173

4.3.6.3 STAT6 and IL-13


Eotaxin is of great interest in studies of asthma and atopy. Elevated levels of eotaxin are routinely
present in asthma patients.174 Several cell types secrete eotaxin in response to IL-13 treatment.
For example, airway epithelial cells induced by IL-13 secrete eotaxin.175 In human skin cells, when
treated with STAT6 siRNA and stimulated with IL-13, eotaxin levels are decreased.164 When
human bronchial epithelial cells are treated with IL-13, STAT6 is activated, suggesting a link
between IL-13, STAT6, and eotaxin, and potentially explaining the mechanism for increased
levels of eosinophils in some forms of asthma.176 In in vitro studies using human airway cell lines,
such as A549, eotaxin family levels (CCL11, CCL24, CCL26) are upregulated in response to IL-4
and IL-13. In this same study, STAT6 siRNA blocked the upregulation of these eotaxins. When
secreted protein levels were examined, only CCL26 was detectable while the others were not. In
this same study, eotaxin gene expression was decreased when primary human bronchial smooth
muscle cells and primary epithelial cells were treated with IL-13 and STAT6 siRNA.177
In a study where transgenic mice that expressed STAT6 under control of a CC10 promoter,
a protein that is specific to airway epithelial cells, were crossed to STAT6 deficient mice, they
selectively expressed STAT6 only in lung epithelium cells. These mice were crossed further with
transgenic IL-13 mice and examined for the requirement of STAT6 in IL-13 induced asthma. While
researchers found that STAT6-deficient mice were protected from airway hyperresponsiveness, they
saw that mice that only expressed STAT6 in the airways still developed airway hyperresponsiveness.178

4.3.5.4 STAT6 and IL-13/IL-17


Like IL-4 and IL-13, IL-17A is increased in asthmatics. IL-17A induces bronchial cells to secrete
cytokines such as IL.-6179 IL-17 can induce the secretion of IL.-13180 In studies that transferred
TH17 cells into both WT and STAT6-deficient mice, IL-13 protein expression was assessed after
OVA challenge. Harvested lungs from both groups showed higher IL-13 protein expression when
compared to PBS controls. When comparing both the OVA-treated groups, the STAT6-deficient
mice displayed higher levels of IL-13 protein expression. Overall, STAT6 negatively regulated
IL-13 protein expression.181
The role of IL-17 on receptor-mediated signaling has been explored. The IL-4Rα chain can link
with the IL-13Rα2 receptor to form the type II IL-4/IL-13 receptor. When this happens, IL-17
can induce IL-13 secretion. Thus, studies have attempted to identify the role of IL-17 on the IL-13
receptor complex.182 Recent studies focusing on IL-13rα2, which has traditionally not been linked to
the development of asthma, have shown that it makes a modest contribution to the development of
airway hyperresponsiveness.183 When IL-13rα2-deficient mice were given IL-17, IL-13 driven airway
hyperresponsiveness was enhanced.180 While IL-13rα1 clearly plays a role in STAT6-mediated
signaling, it appears that IL-13rα2 does not. When non-transgenic mice are compared to IL-13rα2
deficient mice, and STAT6 levels are examined, differences in STAT6 activation are not observed.183

4.3.5.5 STAT6 and Mucus Gene Overproduction


Since one of the hallmark features of asthma is mucus overproduction, studies have defined ways
that STAT6 may, directly or indirectly, play a role in the regulation of mucus genes.127,132,184
58 JAK-STAT Signaling in Diseases

Twelve mucin-related genes are expressed in humans and several of these, such as MUC2 and
MUC4, are sometimes increased in asthmatics when compared to healthy controls.185,186 One of
the more heavily studied genes, MUC5AC, is increased up to 60% higher than normal in
asthmatic patients.185
Intratracheal IL-13 induces MUC5AC in wild type but not STAT6-deficient mice. Consistent
with this, when cells were isolated and then treated with IL-13 for 24 hours, MUC5AC
expression was only slightly increased in STAT6 deficient animals compared to untreated cells,
and to wild type controls.132 Moreover, another gene involved in goblet cell metaplasia, Gob-5,
was unresponsive in cells deficient in STAT6 as compared to wild type controls, suggesting
a role for STAT6 in controlling both genes.132 In a study using STAT6 siRNA to knockdown
STAT6 levels, Muc5ac promoter activation was strongly decreased.187 In another study using
STAT6 siRNA, repression of STAT6 in the presence of IL-13 led to decreased MUC5AC
expression as well as SAM domain-containing prostate-derived Ets factor (SPDEF), which is
a transcription factor that plays a role in goblet cell hyperplasia and mucus secretion.188
Reductions in IL-13-mediated expression of the transmembrane protein 16A, a calcium-
activated chloride channel that is a key regulator of mucus overproduction in airway epithelial
cells, were shown to be a result of STAT6 inhibition. This reduction corresponded to decreased
MUC5AC expression, further highlighting the role of STAT6 in mucin expression.189 In a study
exploring the effects of using a Chinese herb extract, Glycyrrhiza uralensis, on asthma, the
flavonoid 7′4′-dihydroxyflavone was shown to possess anti-inflammatory properties. When this
compound was used on human cell cultures stimulated with PMA, Muc5AC secretion was
decreased compared to untreated controls. This decrease was positively correlated with declining
phosphorylated STAT6.190
The role of STAT6 in controlling mucin genes has additionally been explored in Lyn kinase­
deficient mice. Lyn kinase negatively regulates the progression of asthma.191 In an in vitro study
using human bronchial epithelial cells treated with either IL-4 or IL-13, when Lyn siRNA was
used, STAT6 expression and phosphorylation increased in both conditions, as compared with
untreated cells. When Lyn was overexpressed in their system, STAT6 expression was decreased.
Taken together, the results suggest that Lyn regulates MUC5AC expression via the STAT6
pathway.187

4.3.6 STAT Inhibitors


Theoretically, like JAKinibs, inhibitors of STAT proteins could be promising in immune-related
disorders (see Table 4.2). Efforts in studying STAT inhibitors have been challenged by a lack of
bioavailability and low efficacy and specificity. Some groups have tried to circumvent issues of
specificity by employing oligonucleotide-based STAT inhibitors. One example is the STAT1 decoy
oligonucleotide which, in mouse models, has been shown to reduce lung infiltration and antigen-
specific AHR.192
A STAT3 inhibitor that targets the SH2 domain (C1889) has been used to treat animals
intranasally during allergic asthma model. This treatment significantly reduced TH2 and TH17
development and allergic asthma.193 Currently, some small molecule inhibitors of STAT3 are in
clinical trials to be tested for indications other than asthma.194,195
In mouse models of allergic bronchial asthma, IL-13 administered intranasally leads to induced
phosphorylation of STAT6 as well as the upregulation of the RhoA protein, a protein involved in
smooth muscle contraction. In this same study, when human bronchial smooth muscle cells were
used and treated with a STAT6 inhibitor, leflunomide, STAT6 phosphorylation and the RhoA
upregulation induced by IL13 were abolished.196 In another study, when the STAT6 inhibitor,
AS15174999, was used concurrent with a 1-hour exposure of human bronchial smooth muscle
cells to IL-13, STAT6 phosphorylation was inhibited in a dose-dependent manner. Similarly,
RhoA expression was decreased in these cells. In the same study, when STAT6 siRNA was used,
RhoA expression was partially inhibited.197
Asthma and Allergic Airway Inflammation 59

Another approach taken to find inhibitors of STAT proteins is the use of clinically available
drugs that, through unknown mechanisms, reduce STAT activation. For example, the psycho-
trophic drug pimozide has been shown to reduce the phosphorylation of STAT5198 This drug was
recently used in an ex-vivo study to show that the steroid resistance seen in the ILC2 cells of
asthmatic humans was dependent on TSLP and STAT5 activation.199 Additionally, JAK inhibi­
tors are used to assess the role of STAT proteins in vivo. In particular, the JAK1/3 inhibitor
(R256) can prevent STAT5 activation and TH2 development in culture. Treatment of animals with
R256 can reduce AHR, eosinophilia, and mucus production.200

4.4 Conclusion
Despite some of the challenges, JAK/STAT pathway constituents remain attractive targets for
pharmaceutical intervention. Several intrinsic properties of JAK/STAT signaling highlight the
complexity of this pathway. STAT proteins can be expressed in a cell- and tissue-specific manner.
One cytokine can activate multiple JAK and STAT proteins to varying degrees. STAT proteins
can create homodimers and heterodimers which allow for differential DNA binding.7,201 STAT
proteins may also be able to compete with or compensate for one another and some evidence does
exist supporting this claim.7,202 It has therefore been a challenge to study the effects of inhibiting
STAT proteins in multicellular and multi-organ systems and existing studies that tackle this
question should be interpreted cautiously.
The JAK/STAT pathway is clearly central in the development of asthma. STAT3, STAT5,
and STAT6 are required in multiple cell types for myriad cytokine responses. While the role of
other STAT proteins is less clear, there are likely some situations where they are required in the
spectrum of asthma endotypes that span disease pathologies from purely TH2 to more mixed
TH cell phenotypes. As the use of JAKinibs and more specific STAT inhibitors move more
broadly into the clinic, it will be interesting to see how they affect various patients with asthma
and how personalized medicine can help direct the use of drugs to inhibit these critical
pathways.

REFERENCES
1. McKinley, L. et al. TH17 cells mediate steroid-resistant airway inflammation and airway hyperre­
sponsiveness in mice. J. Immunol. Baltim. Md 1950. 181, 4089–4097 (2008).
2. Lai, T. et al. Interleukin-31 expression and relation to disease severity in human asthma. Sci. Rep. 6,
22835 (2016).
3. Stott, B. et al. Human IL-31 is induced by IL-4 and promotes TH2-driven inflammation. J. Allergy
Clin. Immunol. 132, 446–454.e5 (2013).
4. Chen, M. et al. Complex effects of naturally occurring mutations in the JAK3 pseudokinase domain:
Evidence for Interactions between the kinase and pseudokinase domains. Mol. Cell. Biol. 20, 947–
956 (2000).
5. Lindauer, K., Loerting, T., Liedl, K. R. & Kroemer, R. T. Prediction of the structure of human Janus
kinase 2 (JAK2) comprising the two carboxy-terminal domains reveals a mechanism for
autoregulation. Protein Eng. 14, 27–37 (2001).
6. Williams, N. K. et al. Dissecting specificity in the Janus kinases: The structures of JAK-specific
inhibitors complexed to the JAK1 and JAK2 protein tyrosine kinase domains. J. Mol. Biol. 387,
219–232 (2009).
7. Villarino, A. V., Kanno, Y., & O’Shea, J. J. Mechanisms and consequences of Jak–STAT signaling in
the immune system. Nat. Immunol. 18, 374–384 (2017).
8. Zouein, F. A., Duhé, R. J., & Booz, G. W. JAKs go nuclear: Emerging role of nuclear JAK1 and
JAK2 in gene expression and cell growth. Growth Factors Chur Switz. 29, 245–252 (2011).
9. Frucht, D. M. et al. Unexpected and variable phenotypes in a family with JAK3 deficiency. Genes
Immun. 2, 422–432 (2001).
60 JAK-STAT Signaling in Diseases

10. Karaghiosoff, M. et al. Partial impairment of cytokine responses in Tyk2-deficient mice. Immunity.
13, 549–560 (2000).
11. Minegishi, Y. et al. Human tyrosine kinase 2 deficiency reveals its requisite roles in multiple cytokine
signals involved in innate and acquired immunity. Immunity. 25, 745–755 (2006).
12. Leonard, W. J. & O’Shea, J. J. Jaks and STATs: Biological implications. Annu. Rev. Immunol. 16,
293–322 (1998).
13. Howell, M. D., Fitzsimons, C., & Smith, P. A. JAK/STAT inhibitors and other small molecule
cytokine antagonists for the treatment of allergic disease. Ann. Allergy Asthma Immunol. Off. Publ.
Am. Coll. Allergy Asthma Immunol. 120, 367–375 (2018).
14. Malaviya, R. et al. Treatment of allergic asthma by targeting janus kinase 3-dependent leukotriene
synthesis in mast cells with 4-(3’, 5’-dibromo-4’-hydroxyphenyl)amino-6,7-dimethoxyquinazoline
(WHI-P97). J. Pharmacol. Exp. Ther. 295, 912–926 (2000).
15. Simeone-Penney, M. C. et al. Airway epithelial STAT3 is required for allergic inflammation in
a murine model of asthma. J. Immunol. 178, 6191–6199 (2007).
16. Wiegman, C. H. et al. The selective pan-Janus kinase (JAK) inhibitor VR588 demonstrates potent
anti-inflammatory activity in a murine chronic house dust mite (HDM) model of asthma. in C34.
New Basic Science in Asthma: Allergic Inflammation I. A6435–A6435. (American Thoracic Society,
2015). doi:10.1164/ajrccm-conference.2015.191.1_MeetingAbstracts.A6435.
17. O’Shea, J. J. et al. The JAK-STAT pathway: Impact on human disease and therapeutic intervention.
Annu. Rev. Med. 66, 311–328 (2015).
18. Bergauer, A. et al. IFN-α/IFN-λ responses to respiratory viruses in paediatric asthma. Eur. Respir. J.
49, 1600969 (2017).
19. Jordan, W. J. et al. Human interferon lambda-1 (IFN-λ1/IL-29) modulates the Th1/Th2 response.
Genes Immun. 8, 254–261 (2007).
20. Koltsida, O. et al. IL-28A (IFN-λ2) modulates lung DC function to promote Th1 immune skewing
and suppress allergic airway disease. EMBO Mol. Med. 3, 348–361 (2011).
21. Srinivas, S. et al. Interferon-λ1 (interleukin-29) preferentially down-regulates interleukin-13 over
other T helper type 2 cytokine responses in vitro. Immunology. 125, 492–502 (2008).
22. Shi, Z. O.-Q., Fischer, M. J., Sanctis, G. T. D., Schuyler, M. R., & Tesfaigzi, Y. IFN-γ, but not fas,
mediates reduction of allergen-induced mucous cell metaplasia by inducing apoptosis. J. Immunol.
168, 4764–4771 (2002).
23. Fulkerson, P. C., Zimmermann, N., Hassman, L. M., Finkelman, F. D., & Rothenberg, M. E.
Pulmonary chemokine expression is coordinately regulated by STAT1, STAT6, and IFN-γ.
J. Immunol. 173, 7565–7574 (2004).
24. Sampath, D., Castro, M., Look, D. C., & Holtzman, M. J. Constitutive activation of an epithelial
signal transducer and activator of transcription (STAT) pathway in asthma. J. Clin. Invest. 103,
1353–1361 (1999).
25. Wjst, M., Lichtner, P., Meitinger, T., & Grimbacher, B. STAT3 single-nucleotide polymorphisms and
STAT3 mutations associated with hyper-IgE syndrome are not responsible for increased serum IgE
serum levels in asthma families. Eur. J. Hum. Genet. 17, 352–356 (2009).
26. Zafra, M. P. et al. SOCS3 silencing attenuates eosinophil functions in asthma patients. Int. J. Mol.
Sci. 16, 5434–5451 (2015).
27. Stritesky, G. L. et al. The transcription factor STAT3 is required for T helper 2 cell development.
Immunity. 34, 39–49 (2011).
28. Ulrich, B. J., Verdan, F. F., McKenzie, A. N. J., Kaplan, M. H., & Olson, M. R. STAT3 activation
impairs the stability of Th9 cells. J. Immunol. Baltim. Md 1950. 198, 2302–2309 (2017).
29. Egwuagu, C. E. STAT3 in CD4+ T helper cell differentiation and inflammatory diseases. Cytokine.
47, 149–156 (2009).
30. Yamamoto, K. et al. Stat4, a novel gamma interferon activation site-binding protein expressed in
early myeloid differentiation. Mol. Cell. Biol. 14, 4342–4349 (1994).
31. Nakayamada, S. et al. Early Th1 cell differentiation is marked by a Tfh cell-like transition. Immunity.
35, 919–931 (2011).
32. Schmitt, N. et al. The cytokine TGF-β co-opts signaling via STAT3-STAT4 to promote the
differentiation of human TFH cells. Nat. Immunol. 15, 856–865 (2014).
Asthma and Allergic Airway Inflammation 61

33. Jacobson, N. G. et al. Interleukin 12 signaling in T helper type 1 (Th1) cells involves tyrosine
phosphorylation of signal transducer and activator of transcription (Stat)3 and Stat4. J. Exp. Med.
181, 1755–1762 (1995).
34. Kaplan, M. H., Sun, Y. L., Hoey, T., & Grusby, M. J. Impaired IL-12 responses and enhanced
development of Th2 cells in Stat4-deficient mice. Nature. 382, 174–177 (1996).
35. Kaplan, M. H. & Grusby, M. J. Regulation of T helper cell differentiation by STAT molecules.
J. Leukoc. Biol. 64, 2–5 (1998).
36. Grant, L. et al. Stat4-dependent, T-bet-independent regulation of IL-10 in NK cells. Genes Immun.
9, 316–327 (2008).
37. Glosson-Byers, N. L., Sehra, S., & Kaplan, M. H. STAT4 is required for IL-23 responsiveness in
Th17 memory cells and NKT cells. JAK-STAT 3, e955393 (2014).
38. Xu, J. et al. Stat4 is critical for the balance between Th17 cells and regulatory T cells in colitis.
J. Immunol. Baltim. Md 1950. 186, 6597–6606 (2011).
39. Iida, K. et al. STAT4 is required for IFN-β-induced MCP-1 mRNA expression in murine mast cells.
Int. Arch. Allergy Immunol. 155 Suppl 1, 71–76 (2011).
40. Di Stefano, A. et al. STAT4 activation in smokers and patients with chronic obstructive pulmonary
disease. Eur. Respir. J. 24, 78–85 (2004).
41. Zhang, Y., Li, J., Wang, C., & Zhang, L. Association between the interaction of key genes involved in
effector T-cell pathways and susceptibility to developallergic rhinitis: A population-based
case-control association study. PLoS One. 10, e0131248 (2015).
42. Li, X. et al. Genome-wide association study identifies TH1 pathway genes associated with lung
function in asthmatic patients. J. Allergy Clin. Immunol. 132, 313–320.e15 (2013).
43. Nakamura, Y. et al. Therapeutic implication of genetic variants of IL13 and STAT4 in airway
remodelling with bronchial asthma. Clin. Exp. Allergy J. Br. Soc. Allergy Clin. Immunol. 46, 1152–
1161 (2016).
44. Nguyen, K. B. et al. Critical role for STAT4 activation by type 1 interferons in the interferon-gamma
response to viral infection. Science. 297, 2063–2066 (2002).
45. Wurtz, O., Bajénoff, M., & Guerder, S. IL-4-mediated inhibition of IFN-gamma production by
CD4+ T cells proceeds by several developmentally regulated mechanisms. Int. Immunol. 16, 501–
508 (2004).
46. Kaplan, M. H., Schindler, U., Smiley, S. T., & Grusby, M. J. Stat6 is required for mediating responses
to IL-4 and for development of Th2 cells. Immunity. 4, 313–319 (1996).
47. Ferber, I. A. et al. GATA-3 significantly downregulates IFN-gamma production from developing
Th1 cells in addition to inducing IL-4 and IL-5 levels. Clin. Immunol. Orlando Fla. 91, 134–144
(1999).
48. Ouyang, W. et al. Inhibition of Th1 development mediated by GATA-3 through an
IL-4-independent mechanism. Immunity. 9, 745–755 (1998).
49. Chen, Y. et al. IFN-γ-expressing Th17 cells are required for development of severe ocular surface
autoimmunity. J. Immunol. Baltim. Md 1950. 199, 1163–1169 (2017).
50. Annunziato, F. et al. Phenotypic and functional features of human Th17 cells. J. Exp. Med. 204,
1849–1861 (2007).
51. Kotake, S., Yago, T., Kobashigawa, T., & Nanke, Y. The plasticity of Th17 cells in the pathogenesis
of rheumatoid arthritis. J. Clin. Med. 6, E67 (2017).
52. Glosson-Byers, N. L. et al. Th17 cells demonstrate stable cytokine production in a proallergic
environment. J. Immunol. Baltim. Md 1950. 193, 2631–2640 (2014).
53. Mathur, A. N. et al. T-bet is a critical determinant in the instability of the IL-17-secreting T-helper
phenotype. Blood. 108, 1595–1601 (2006).
54. Lee, Y. K. et al. Late developmental plasticity in the T helper 17 lineage. Immunity. 30, 92–107
(2009).
55. Gavett, S. H. et al. Interleukin 12 inhibits antigen-induced airway hyperresponsiveness, inflamma­
tion, and Th2 cytokine expression in mice. J. Exp. Med. 182, 1527–1536 (1995).
56. Schwarze, J. et al. Local treatment with IL-12 is an effective inhibitor of airway hyperresponsiveness
and lung eosinophilia after airway challenge in sensitized mice. J. Allergy Clin. Immunol. 102, 86–93
(1998).
62 JAK-STAT Signaling in Diseases

57. Hofstra, C. L. et al. Prevention of Th2-like cell responses by coadministration of IL-12 and IL-18 is
associated with inhibition of antigen-induced airway hyperresponsiveness, eosinophilia, and serum
IgE levels. J. Immunol. Baltim. Md 1950. 161, 5054–5060 (1998).
58. Iwamoto, I., Kumano, K., Kasai, M., Kurasawa, K., & Nakao, A. Interleukin-12 prevents
antigen-induced eosinophil recruitment into mouse airways. Am. J. Respir. Crit. Care Med. 154,
1257–1260 (1996).
59. Matsubara, S. et al. IL-2 and IL-18 attenuation of airway hyperresponsiveness requires STAT4,
IFN-γ, and natural killer cells. Am. J. Respir. Cell Mol. Biol. 36, 324–332 (2007).
60. Meuronen, A. et al. Attenuated expression of tenascin-c in ovalbumin-challenged STAT4−/− mice.
Respir. Res. 12, 2 (2011).
61. Raman, K., Kaplan, M. H., Hogaboam, C. M., Berlin, A., & Lukacs, N. W. STAT4 signal pathways
regulate inflammation and airway physiology changes in allergic airway inflammation locally via
alteration of chemokines. J. Immunol. 170, 3859–3865 (2003).
62. Thibodeaux, D. K. et al. Autocrine regulation of IL-12 receptor expression is independent of
secondary IFN-γ secretion and not restricted to T and NK cells. J. Immunol. 163, 5257–5264 (1999).
63. Lawless, V. A. et al. Stat4 regulates multiple components of IFN-gamma-inducing signaling
pathways. J. Immunol. Baltim. Md 1950. 165, 6803–6808 (2000).
64. Sur, S. et al. Long term prevention of allergic lung inflammation in a mouse model of asthma by
CpG oligodeoxynucleotides. J. Immunol. 162, 6284–6293 (1999).
65. Kim, Y.-K. et al. Airway exposure levels of lipopolysaccharide determine type 1 versus type 2
experimental asthma. J. Immunol. Baltim. Md 1950. 178, 5375–5382 (2007).
66. Kim, Y.-S. et al. IL-12-STAT4-IFN-γ axis is a key downstream pathway in the development of IL-
13-mediated asthma phenotypes in a Th2 type asthma model. Exp. Mol. Med. 42, 533–546 (2010).
67. Zhang, Y.-L. et al. Peripheral blood MDSCs, IL-10 and IL-12 in children with asthma and their
importance in asthma development. PLoS One. 8, e63775 (2013).
68. Chen, T. et al. Association of single nucleotide polymorphisms in interleukin 12 (IL-12A and -B)
with asthma in a Chinese population. Hum. Immunol. 72, 603–606 (2011).
69. Cui, A.-H., Zhao, J., Liu, S.-X., & Hao, Y.-S. Associations of IL-4, IL-6, and IL-12 levels in
peripheral blood with lung function, cellular immune function, and quality of life in children with
moderate-to-severe asthma. Medicine (Baltimore). 96, e6265 (2017).
70. Schwantes, E. A. et al. Interferon gene expression in sputum cells correlates with the Asthma Index
Score during virus-induced exacerbations. Clin. Exp. Allergy J. Br. Soc. Allergy Clin. Immunol. 44,
813–821 (2014).
71. Brooks, G. D., Buchta, K. A., Swenson, C. A., Gern, J. E., & Busse, W. W. Rhinovirus-induced
interferon-gamma and airway responsiveness in asthma. Am. J. Respir. Crit. Care Med. 168, 1091–
1094 (2003).
72. Smart, J. M., Horak, E., Kemp, A. S., Robertson, C. F., & Tang, M. L. K. Polyclonal and
allergen-induced cytokine responses in adults with asthma: Resolution of asthma is associated with
normalization of IFN-gamma responses. J. Allergy Clin. Immunol. 110, 450–456 (2002).
73. Noma, T. et al. Pattern of cytokine production by T cells from adolescents with asthma in remission,
after stimulation with dermatophagoides farinae antigen. Pediatr. Res. 38, 187–193 (1995).
74. Leonard, C., Tormey, V., Burke, C., & Poulter, L. W. Allergen-induced cytokine production in atopic
disease and its relationship to disease severity. Am. J. Respir. Cell Mol. Biol. 17, 368–375 (1997).
75. Smith, N. L. D. & Denning, D. W. Clinical implications of interferon-γ genetic and epigenetic
variants. Immunology. 143, 499–511 (2014).
76. Litonjua, A. A. et al. Serum interferon-gamma is associated with longitudinal decline in lung
function among asthmatic patients: The Normative Aging Study. Ann. Allergy Asthma Immunol.
Off. Publ. Am. Coll. Allergy Asthma Immunol. 90, 422–428 (2003).
77. Runyon, R. S. et al. Asthma discordance in twins is linked to epigenetic modifications of T cells.
PLoS One. 7, e48796 (2012).
78. Ouyang, B. et al. Interferon-γ promoter is hypermethylated in blood DNA from workers with
confirmed diisocyanate asthma. Toxicol. Sci. 133, 218–224 (2013).
79. Pykäläinen, M. et al. Association analysis of common variants of STAT6, GATA3, and STAT4 to
asthma and high serum IgE phenotypes. J. Allergy Clin. Immunol. 115, 80–87 (2005).
Asthma and Allergic Airway Inflammation 63

80. Singh, D. P., Bagam, P., Sahoo, M. K., & Batra, S. Immune-related gene polymorphisms in
pulmonary diseases. Toxicology. 383, 24–39 (2017).
81. Park, B. L. et al. Association analysis of signal transducer and activator of transcription 4 (STAT4)
polymorphisms with asthma. J. Hum. Genet. 50, 133–138 (2005).
82. Loisel, D. A. et al. Genetic associations with viral respiratory illnesses and asthma control in
children. Clin. Exp. Allergy J. Br. Soc. Allergy Clin. Immunol. 46, 112–124 (2016).
83. Li, Y., Wu, B., Xiong, H., Zhu, C., & Zhang, L. Polymorphisms of STAT-6, STAT-4 and
IFN-gamma genes and the risk of asthma in Chinese population. Respir. Med. 101, 1977–1981
(2007).
84. Caslin, H. L. et al. Controlling mast cell activation and homeostasis: Work influenced by Bill Paul
that continues today. Front. Immunol. 9, 868 (2018).
85. Kanai, T., Jenks, J., & Nadeau, K. C. The STAT5b pathway defect and autoimmunity. Front.
Immunol. 3, 234 (2012).
86. Saeedfar, K., Behmanesh, M., Mortaz, E., & Masjedi, M. R. The expression of STAT3 and STAT5A
genes in severe refractory asthma. Tanaffos. 16, 1–8 (2017).
87. Barnstein, B. O. et al. Stat5 expression is required for IgE-mediated mast cell function. J. Immunol.
177, 3421–3426 (2006).
88. Schmitt, E. et al. IL-9 production of naive CD4+ T cells depends on IL-2, is synergistically enhanced
by a combination of TGF-beta and IL-4, and is inhibited by IFN-gamma. J. Immunol. 153, 3989–
3996 (1994).
89. Papathanassiou, E., Loukides, S., & Bakakos, P. Severe asthma: Anti-IgE or anti-IL-5? Eur. Clin.
Respir. J. 3, 31813 (2016).
90. Liao, W. et al. Opposing actions of IL-2 and IL-21 on Th9 differentiation correlate with their
differential regulation of BCL6 expression. Proc. Natl. Acad. Sci. 111, 3508–3513 (2014).
91. Olson, M. R., Verdan, F. F., Hufford, M. M., Dent, A. L., & Kaplan, M. H. STAT3 impairs STAT5
activation in the development of IL-9–secreting T cells. J. Immunol. 196, 3297–3304 (2016).
92. Li, G., Liu, Z., Ran, P., Qiu, J., & Zhong, N. Activation of signal transducer and activator of
transcription 5 (STAT5) in splenocyte proliferation of asthma mice induced by ovalbumin. Cell.
Mol. Immunol. 1, 471–474 (2004).
93. Hamid, Q. et al. Expression of mRNA for interleukin-5 in mucosal bronchial biopsies from asthma.
J. Clin. Invest. 87, 1541–1546 (1991).
94. Haldar, P. et al. Mepolizumab and exacerbations of refractory eosinophilic asthma. N. Engl. J. Med.
360, 973–984 (2009).
95. Ortega, H. G. et al. Mepolizumab treatment in patients with severe eosinophilic asthma. N. Engl.
J. Med. 371, 1198–1207 (2014).
96. Bel, E. H. et al. Oral glucocorticoid-sparing effect of mepolizumab in eosinophilic asthma. N. Engl.
J. Med. 371, 1189–1197 (2014).
97. Kopf, M. et al. IL-5-deficient mice have a developmental defect in CD5+ B-1 cells and lack
eosinophilia but have normal antibody and cytotoxic T cell responses. Immunity. 4, 15–24 (1996).
98. Foster, P. S., Hogan, S. P., Ramsay, A. J., Matthaei, K. I., & Young, I. G. Interleukin 5 deficiency
abolishes eosinophilia, airways hyperreactivity, and lung damage in a mouse asthma model. J. Exp.
Med. 183, 195–201 (1996).
99. Kagami, S. et al. Both Stat5a and Stat5b are required for antigen-induced eosinophil and T-cell
recruitment into the tissue. Blood. 95, 1370–1377 (2000).
100. Kagami, S. et al. Stat5a regulates T helper cell differentiation by several distinct mechanisms. Blood.
97, 2358–2365 (2001).
101. Zhu, J., Cote-Sierra, J., Guo, L., & Paul, W. E. Stat5 activation plays a critical role in Th2
differentiation. Immunity. 19, 739–748 (2003).
102. Takatori, H. et al. Indispensable role of Stat5a in Stat6-independent Th2 cell differentiation and
allergic airway inflammation. J. Immunol. 174, 3734–3740 (2005).
103. Rochman, Y. et al. Thymic stromal lymphopoietin-mediated STAT5 phosphorylation via kinases
JAK1 and JAK2 reveals a key difference from IL-7–induced signaling. Proc. Natl. Acad. Sci.
U. S. A. 107, 19455–19460 (2010).
104. Kashyap, M., Rochman, Y., Spolski, R., Samsel, L., & Leonard, W. J. Thymic stromal lymphopoie­
tin is produced by dendritic cells. J. Immunol. Baltim. Md 1950. 187, 1207–1211 (2011).
64 JAK-STAT Signaling in Diseases

105. Zhou, B. et al. Thymic stromal lymphopoietin as a key initiator of allergic airway inflammation in
mice. Nat. Immunol. 6, 1047–1053 (2005).
106. Al-Shami, A., Spolski, R., Kelly, J., Keane-Myers, A., & Leonard, W. J. A role for TSLP in the
development of inflammation in an asthma model. J. Exp. Med. 202, 829–839 (2005).
107. Nguyen, K. D., Vanichsarn, C., & Nadeau, K. C. TSLP directly impairs pulmonary Treg function:
Association with aberrant tolerogenic immunity in asthmatic airway. Allergy Asthma Clin. Immunol.
Off. J. Can. Soc. Allergy Clin. Immunol. 6, 4 (2010).
108. Lei, L., Zhang, Y., Yao, W., Kaplan, M. H., & Zhou, B. TSLP interferes with airway tolerance by
suppressing the generation of antigen-specific regulatory T cells. J. Immunol. Baltim. Md 1950. 186,
2254–2261 (2011).
109. Quelle, F. W. et al. Cloning of murine Stat6 and human Stat6, Stat proteins that are tyrosine
phosphorylated in responses to IL-4 and IL-3 but are not required for mitogenesis. Mol. Cell. Biol.
15, 3336–3343 (1995).
110. Hou, J. et al. An interleukin-4-induced transcription factor: IL-4 Stat. Science. 265, 1701–1706
(1994).
111. Takeda, K., Kamanaka, M., Tanaka, T., Kishimoto, T., & Akira, S. Impaired IL-13-mediated
functions of macrophages in STAT6-deficient mice. J. Immunol. Baltim. Md 1950. 157, 3220–3222
(1996).
112. Hershey, G. K. K. IL-13 receptors and signaling pathways: An evolving web. J. Allergy Clin.
Immunol. 111, 677–690; quiz 691 (2003).
113. Masuda, A. et al. Interleukin-15 induces rapid tyrosine phosphorylation of STAT6 and the
expression of interleukin-4 in mouse mast cells. J. Biol. Chem. 275, 29331–29337 (2000).
114. Masuda, A., Matsuguchi, T., Yamaki, K., Hayakawa, T., & Yoshikai, Y. Interleukin-15 prevents
mouse mast cell apoptosis through STAT6-mediated Bcl-xL expression. J. Biol. Chem. 276, 26107–
26113 (2001).
115. Hebenstreit, D., Wirnsberger, G., Horejs-Hoeck, J., & Duschl, A. Signaling mechanisms, interaction
partners, and target genes of STAT6. Cytokine Growth Factor Rev. 17, 173–188 (2006).
116. Maier, E., Duschl, A., & Horejs-Hoeck, J. STAT6-dependent and -independent mechanisms in Th2
polarization. Eur. J. Immunol. 42, 2827–2833 (2012).
117. Asnagli, H., Afkarian, M., & Murphy, K. M. Cutting edge: Identification of an alternative GATA-3
promoter directing tissue-specific gene expression in mouse and human. J. Immunol. Baltim. Md
1950. 168, 4268–4271 (2002).
118. Stütz, A. M. et al. The Th2 cell cytokines IL-4 and IL-13 regulate found in inflammatory zone 1/
resistin-like molecule alpha gene expression by a STAT6 and CCAAT/enhancer-binding
protein-dependent mechanism. J. Immunol. Baltim. Md 1950. 170, 1789–1796 (2003).
119. Hoeck, J. & Woisetschläger, M. Activation of eotaxin-3/CCLl26 gene expression in human dermal
fibroblasts is mediated by STAT6. J. Immunol. Baltim. Md 1950. 167, 3216–3222 (2001).
120. Matsukura, S. et al. Activation of eotaxin gene transcription by NF-kappa B and STAT6 in human
airway epithelial cells. J. Immunol. Baltim. Md 1950. 163, 6876–6883 (1999).
121. Wirnsberger, G., Hebenstreit, D., Posselt, G., Horejs-Hoeck, J., & Duschl, A. IL-4 induces expres­
sion of TARC/CCL17 via two STAT6 binding sites. Eur. J. Immunol. 36, 1882–1891 (2006).
122. Hoeck, J. & Woisetschläger, M. STAT6 mediates eotaxin-1 expression in IL-4 or TNF-alpha­
induced fibroblasts. J. Immunol. Baltim. Md 1950. 166, 4507–4515 (2001).
123. Maier, E., Wirnsberger, G., Horejs-Hoeck, J., Duschl, A., & Hebenstreit, D. Identification of a distal
tandem STAT6 element within the CCL17 locus. Hum. Immunol. 68, 986–992 (2007).
124. Goenka, S. & Kaplan, M. H. Transcriptional regulation by STAT6. Immunol. Res. 50, 87–96 (2011).
125. Nguyen, V. T. & Benveniste, E. N. IL-4-activated STAT-6 inhibits IFN-gamma-induced CD40 gene
expression in macrophages/microglia. J. Immunol. Baltim. Md 1950. 165, 6235–6243 (2000).
126. Bennett, B. L., Cruz, R., Lacson, R. G., & Manning, A. M. Interleukin-4 suppression of tumor
necrosis factor α-stimulated E-selectin gene transcription is mediated by STAT6 antagonism of NF­
κB. J. Biol. Chem. 272, 10212–10219 (1997).
127. Holgate, S. T. Pathogenesis of asthma. Clin. Exp. Allergy J. Br. Soc. Allergy Clin. Immunol. 38, 872–
897 (2008).
128. Umetsu, D. T. & DeKruyff, R. H. A role for natural killer T cells in asthma. Nat. Rev. Immunol. 6,
953–958 (2006).
Asthma and Allergic Airway Inflammation 65

129. Akbari, O. et al. CD4+ invariant T-cell-receptor+ natural killer T cells in bronchial asthma. N. Engl.
J. Med. 354, 1117–1129 (2006).
130. Bradding, P., Walls, A. F., & Holgate, S. T. The role of the mast cell in the pathophysiology of
asthma. J. Allergy Clin. Immunol. 117, 1277–1284 (2006).
131. Brightling, C. E. et al. Interleukin-4 and -13 expression is co-localized to mast cells within the airway
smooth muscle in asthma. Clin. Exp. Allergy J. Br. Soc. Allergy Clin. Immunol. 33, 1711–1716
(2003).
132. Thai, P., Chen, Y., Dolganov, G., & Wu, R. Differential regulation of MUC5AC/Muc5ac and
hCLCA-1/mGob-5 expression in airway epithelium. Am. J. Respir. Cell Mol. Biol. 33, 523–530
(2005).
133. Venkayya, R. et al. The Th2 lymphocyte products IL-4 and IL-13 rapidly induce airway hyperre­
sponsiveness through direct effects on resident airway cells. Am. J. Respir. Cell Mol. Biol. 26, 202–
208 (2002).
134. Huo, Y. & Zhang, H.-Y. Genetic mechanisms of asthma and the implications for drug repositioning.
Genes 9, E237 (2018).
135. Goswami, R. et al. STAT6-dependent regulation of Th9 development. J. Immunol. Baltim. Md 1950.
188, 968–975 (2012).
136. Zhu, J., Guo, L., Watson, C. J., Hu-Li, J., & Paul, W. E. Stat6 is necessary and sufficient for IL-4’s
role in Th2 differentiation and cell expansion. J. Immunol. 166, 7276–7281 (2001).
137. Summers, S. A. et al. Signal transducer and activation of transcription 6 (STAT6) regulates T helper
type 1 (Th1) and Th17 nephritogenic immunity in experimental crescentic glomerulonephritis. Clin.
Exp. Immunol. 166, 227–234 (2011).
138. Doherty, T. A. et al. Alternaria induces STAT6-dependent acute airway eosinophilia and epithelial
FIZZ1 expression that promotes airway fibrosis and epithelial thickness. J. Immunol. Baltim. Md
1950. 188, 2622–2629 (2012).
139. Krishnamurthy, P. & Kaplan, M. H. STAT6 and PARP family members in the development of T
cell-dependent allergic inflammation. Immune Netw. 16, 201–210 (2016).
140. Antczak, A. et al. Analysis of changes in expression of IL-4/IL-13/STAT6 pathway and correlation
with the selected clinical parameters in patients with atopic asthma. Int. J. Immunopathol. Pharma­
col. 29, 195–204 (2016).
141. Gao, P. S. et al. Variants of STAT6 (signal transducer and activator of transcription 6) in atopic
asthma. J. Med. Genet. 37, 380–382 (2000).
142. Al-Muhsen, S. et al. Association of the STAT-6 rs324011 (C2892T) variant but not rs324015
(G2964A), with atopic asthma in a Saudi Arabian population. Hum. Immunol. 75, 791–795 (2014).
143. Nagarkatti, R., B-Rao, C., Vijayan, V., Sharma, S. K., & Ghosh, B. Signal transducer and activator
of transcription 6 haplotypes and asthma in the Indian population. Am. J. Respir. Cell Mol. Biol. 31,
317–321 (2004).
144. Godava, M., Vrtel, R., & Vodicka, R. STAT6 - polymorphisms, haplotypes and epistasis in relation
to atopy and asthma. Biomed. Pap. Med. Fac. Univ. Palacky Olomouc Czechoslov. 157, 172–180
(2013).
145. Eifan, A. O. et al. Reduced T-bet in addition to enhanced STAT6 and GATA3 expressing T cells
contribute to human allergen-induced late responses. Clin. Exp. Allergy J. Br. Soc. Allergy Clin.
Immunol. 42, 891–900 (2012).
146. Yu, Q. N. et al. ILC2 frequency and activity are inhibited by glucocorticoid treatment via STAT
pathway in patients with asthma. Allergy. 73, 1860–1870 (2018).
147. Taha, R., Hamid, Q., Cameron, L., & Olivenstein, R. T helper type 2 cytokine receptors and
associated transcription factors GATA-3, c-MAF, and signal transducer and activator of transcrip­
tion factor-6 in induced sputum of atopic asthmatic patients. Chest. 123, 2074–2082 (2003).
148. Stephenson, L., Johns, M. H., Woodward, E., Mora, A. L., & Boothby, M. An IL-4R alpha allelic
variant, I50, acts as a gain-of-function variant relative to V50 for Stat6, but not Th2 differentiation.
J. Immunol. Baltim. Md 1950. 173, 4523–4528 (2004).
149. Risma, K. A. et al. V75R576 IL-4 receptor α is associated with allergic asthma and enhanced IL-4
receptor function. J. Immunol. 169, 1604–1610 (2002).
150. Khoo, S.-K. et al. Associations of a novel IL4RA polymorphism, Ala57Thr, in Greenlander Inuit.
J. Allergy Clin. Immunol. 118, 627–634 (2006).
66 JAK-STAT Signaling in Diseases

151. Kruse, S. et al. The polymorphisms S503P and Q576R in the interleukin-4 receptor alpha gene are
associated with atopy and influence the signal transduction. Immunology. 96, 365–371 (1999).
152. Mitsuyasu, H. et al. Ile50Val variant of IL4R alpha upregulates IgE synthesis and associates with
atopic asthma. Nat. Genet. 19, 119–120 (1998).
153. Ford, A. Q., Heller, N. M., Stephenson, L., Boothby, M. R., & Keegan, A. D. An atopy-associated
polymorphism in the ectodomain of the IL-4Rα chain (V50) regulates the persistence of STAT6
phosphorylation. J. Immunol. Baltim. Md 1950. 183, 1607–1616 (2009).
154. Qian, X., Gao, Y., Ye, X., & Lu, M. Association of STAT6 variants with asthma risk: A systematic
review and meta-analysis. Hum. Immunol. 75, 847–853 (2014).
155. Kavalar, M. S. et al. Association of ORMDL3, STAT6 and TBXA2R gene polymorphisms with
asthma. Int. J. Immunogenet. 39, 20–25 (2012).
156. Demenais, F. et al. Multiancestry association study identifies new asthma risk loci that colocalize
with immune-cell enhancer marks. Nat. Genet. 50, 42–53 (2018).
157. Xie, T. et al. Rho-kinase inhibitor fasudil reduces allergic airway inflammation and mucus hyperse­
cretion by regulating STAT6 and NFκB. Clin. Exp. Allergy J. Br. Soc. Allergy Clin. Immunol. 45,
1812–1822 (2015).
158. Fu, L. S. et al. Heparin protects BALB/c mice from mite-induced airway allergic inflammation. Int.
J. Immunopathol. Pharmacol. 26, 349–359 (2013).
159. Ahmed, T., Smith, G., Vlahov, I., & Abraham, W. M. Inhibition of allergic airway responses by
heparin derived oligosaccharides: Identification of a tetrasaccharide sequence. Respir. Res. 13, 6
(2012).
160. Maarsingh, H., de Boer, J., Kauffman, H. F., Zaagsma, J., & Meurs, H. Heparin normalizes
allergen-induced nitric oxide deficiency and airway hyperresponsiveness. Br. J. Pharmacol. 142,
1293–1299 (2004).
161. Ogawa, T., Shimizu, S., & Shimizu, T. The effect of heparin on antigen-induced mucus hypersecre­
tion in the nasal epithelium of sensitized rats. Allergol. Int. Off. J. Jpn. Soc. Allergol. 62, 77–83
(2013).
162. Bianco, S., Vaghi, A., Robuschi, M., & Pasargiklian, M. Prevention of exercise-induced broncho­
constriction by inhaled frusemide. Lancet Lond. Engl. 2, 252–255 (1988).
163. Ghonim, M. A. et al. Sulfated non-anticoagulant heparin blocks Th2-induced asthma by modulat­
ing the IL-4/signal transducer and activator of transcription 6/Janus kinase 1 pathway. J. Transl.
Med. 16, 243 (2018).
164. Darcan-Nicolaisen, Y. et al. Small interfering RNA against transcription factor STAT6 inhibits
allergic airway inflammation and hyperreactivity in mice. J. Immunol. Baltim. Md 1950. 182, 7501–
7508 (2009).
165. Hosoya, K. et al. Gene silencing of STAT6 with siRNA ameliorates contact hypersensitivity and
allergic rhinitis. Allergy. 66, 124–131 (2011).
166. Blease, K. et al. Stat6-deficient mice develop airway hyperresponsiveness and peribronchial fibrosis
during chronic fungal asthma. Am. J. Pathol. 160, 481–490 (2002).
167. Trifilieff, A., El-Hasim, A., Corteling, R., & Owen, C. E. Abrogation of lung inflammation in
sensitized Stat6-deficient mice is dependent on the allergen inhalation procedure. Br. J. Pharmacol.
130, 1581–1588 (2000).
168. Goebeler, M. et al. Interleukin-13 selectively induces monocyte chemoattractant protein-1 synthesis
and secretion by human endothelial cells. Involvement of IL-4R alpha and Stat6 phosphorylation.
Immunology. 91, 450–457 (1997).
169. Zhang, S., Lukacs, N. W., Lawless, V. A., Kunkel, S. L., & Kaplan, M. H. Cutting edge: Differential
expression of chemokines in Th1 and Th2 cells is dependent on Stat6 but not Stat4. J. Immunol.
Baltim. Md 1950. 165, 10–14 (2000).
170. Ohga, K. et al. YM-341619 suppresses the differentiation of spleen T cells into Th2 cells in vitro,
eosinophilia, and airway hyperresponsiveness in rat allergic models. Eur. J. Pharmacol. 590, 409–416
(2008).
171. Liu, Z. et al. Boswellic acid attenuates asthma phenotypes by downregulation of GATA3 via
pSTAT6 inhibition in a murine model of asthma. Int. J. Clin. Exp. Pathol. 8, 236–243 (2015).
172. Chen, B. L. et al. Tetrahydrocurcumin, a major metabolite of curcumin, ameliorates allergic airway
inflammation by attenuating Th2 response and suppressing the IL-4Rα-Jak1-STAT6 and Jagged1/
Asthma and Allergic Airway Inflammation 67

Jagged2 -Notch1/Notch2 pathways in asthmatic mice. Clin. Exp. Allergy J. Br. Soc. Allergy Clin.
Immunol. (2018). doi:10.1111/cea.13258.
173. Kim, S.-H., Hong, J.-H., & Lee, Y.-C. Ursolic acid, a potential PPARγ agonist, suppresses
ovalbumin-induced airway inflammation and Penh by down-regulating IL-5, IL-13, and IL-17 in
a mouse model of allergic asthma. Eur. J. Pharmacol. 701, 131–143 (2013).
174. Lamkhioued, B. et al. Increased expression of eotaxin in bronchoalveolar lavage and airways of
asthmatics contributes to the chemotaxis of eosinophils to the site of inflammation. J. Immunol.
Baltim. Md 1950. 159, 4593–4601 (1997).
175. Li, L. et al. Effects of Th2 cytokines on chemokine expression in the lung: IL-13 potently
induces eotaxin expression by airway epithelial cells. J. Immunol. Baltim. Md 1950. 162, 2477–
2487 (1999).
176. Matsukura, S. et al. Interleukin-13 upregulates eotaxin expression in airway epithelial cells by a
STAT6-dependent mechanism. Am. J. Respir. Cell Mol. Biol. 24, 755–761 (2001).
177. Walker, W., Healey, G. D., & Hopkin, J. M. RNA interference of STAT6 rapidly attenuates ongoing
interleukin-13-mediated events in lung epithelial cells. Immunology. 127, 256–266 (2009).
178. Kuperman, D. A. et al. Direct effects of interleukin-13 on epithelial cells cause airway hyperreactiv­
ity and mucus overproduction in asthma. Nat. Med. 8, 885–889 (2002).
179. Molet, S. et al. IL-17 is increased in asthmatic airways and induces human bronchial fibroblasts to
produce cytokines. J. Allergy Clin. Immunol. 108, 430–438 (2001).
180. Hall, S. L. et al. IL-17A enhances IL-13 activity by enhancing IL-13-induced signal transducer and
activator of transcription 6 activation. J. Allergy Clin. Immunol. 139, 462–471.e14 (2017).
181. Newcomb, D. C. et al. IL-17A induces signal transducers and activators of transcription-6-independent
airway mucous cell metaplasia. Am. J. Respir. Cell Mol. Biol. 48, 711–716 (2013).
182. Walford, H. H. & Doherty, T. A. STAT6 and lung inflammation. JAK-STAT. 2, e25301 (2013).
183. Chen, W. et al. IL-13 receptor alpha 2 contributes to development of experimental allergic asthma.
J. Allergy Clin. Immunol. 132, 951–958 (2013).
184. Cohn, L. Mucus in chronic airway diseases: Sorting out the sticky details. J. Clin. Invest. 116, 306–
308 (2006).
185. Fahy, J. V. Goblet cell and mucin gene abnormalities in asthma. Chest. 122, 320S–326S (2002).
186. Rose, M. C. Mucins: Structure, function, and role in pulmonary diseases. Am. J. Physiol. 263, L413–
429 (1992).
187. Wang, X. et al. Lyn regulates mucus secretion and MUC5AC via the STAT6 signaling pathway
during allergic airway inflammation. Sci. Rep. 7, 42675 (2017).
188. Yu, H., Li, Q., Kolosov, V. P., Perelman, J. M., & Zhou, X. Interleukin-13 induces mucin 5AC
production involving STAT6/SPDEF in human airway epithelial cells. Cell Commun. Adhes. 17, 83–
92 (2010).
189. Qin, Y. et al. Interleukin-13 stimulates MUC5AC expression via a STAT6-TMEM16A-ERK1/2
pathway in human airway epithelial cells. Int. Immunopharmacol. 40, 106–114 (2016).
190. Liu, C. et al. The flavonoid 7,4’-dihydroxyflavone inhibits MUC5AC gene expression, production,
and secretion via regulation of NF-κB, STAT6, and HDAC2. Phytother. Res. PTR. 29, 925–932
(2015).
191. Beavitt, S.-J. E. et al. Lyn-deficient mice develop severe, persistent asthma: Lyn is a critical negative
regulator of Th2 immunity. J. Immunol. Baltim. Md 1950. 175, 1867–1875 (2005).
192. Quarcoo, D. et al. Inhibition of signal transducer and activator of transcription 1 attenuates
allergen-induced airway inflammation and hyperreactivity. J. Allergy Clin. Immunol. 114, 288–295
(2004).
193. Gavino, A. C., Nahmod, K., Bharadwaj, U., Makedonas, G., & Tweardy, D. J. STAT3 inhibition
prevents lung inflammation, remodeling, and accumulation of Th2 and Th17 cells in a murine
asthma model. Allergy. 71, 1684–1692 (2016).
194. OPB-51602 in locally advanced nasopharyngeal carcinoma prior to definitive chemoradiotherapy ­
Full text view - ClinicalTrials.gov. Available at: https://clinicaltrials.gov/ct2/show/NCT02058017.
(Accessed: 5th October 2018).
195. Phase I/II study of OPB-31121 in patients with progressive hepatocellular carcinoma - Full text
view - ClinicalTrials.gov. Available at: https://clinicaltrials.gov/ct2/show/NCT01406574. (Accessed:
5th October 2018).
68 JAK-STAT Signaling in Diseases

196. Chiba, Y. et al. Interleukin-13 augments bronchial smooth muscle contractility with an
up-regulation of RhoA protein. Am. J. Respir. Cell Mol. Biol. 40, 159–167 (2009).
197. Goto, K. et al. The proximal STAT6 and NF-kappaB sites are responsible for IL-13- and
TNF-alpha-induced RhoA transcriptions in human bronchial smooth muscle cells. Pharmacol.
Res. 61, 466–472 (2010).
198. Nelson, E. A. et al. The STAT5 inhibitor pimozide displays efficacy in models of acute myelogenous
leukemia driven by FLT3 mutations. Genes Cancer. 3, 503–511 (2012).
199. Liu, S. et al. Steroid resistance of airway type 2 innate lymphoid cells from patients with severe
asthma: The role of thymic stromal lymphopoietin. J. Allergy Clin. Immunol. 141, 257–268.e6 (2018).
200. Ashino, S. et al. JAK1/3 signaling pathways are key initiators of TH2 differentiation and lung
allergic responses. J. Allergy Clin. Immunol. 133, 1162–1174.e4 (2014).
201. Delgoffe, G. M. & Vignali, D. A. A. STAT heterodimers in immunity. JAK-STAT 2, e23060 (2013).
202. Yang, X.-P. et al. Opposing regulation of the locus encoding IL-17 through direct, reciprocal actions
of STAT3 and STAT5. Nat. Immunol. 12, 247–254 (2011).
5
Role of JAK-STAT Signaling in Atopic Dermatitis

Radomir M. Slominski
Department of Dermatology
Indiana University School of Medicine
Indianapolis, Indiana

Matthew J. Turner
Department of Dermatology
Indiana University School of Medicine
Indianapolis, Indiana
Richard L. Roudebush VA Medical Center
Indianapolis, Indiana

5.1 Introduction
Atopic dermatitis (AD) is one of the most common inflammatory diseases known to humankind.
Severe itch, lost sleep, and secondary skin infections are major contributors to the morbidity of this
disease (Holm et al. 2006; Ong and Leung 2010). Allergic skin inflammation and associated impaired
skin barrier function are key drivers of AD pathogenesis, and Th2 (and other) cytokines are
implicated in initiation and preservation of skin inflammation in AD. Targeted therapeutics have
recently entered the market for treatment of AD, including monoclonal anti-human IL-4Rα antibody,
dupilumab, which was FDA approved in 2017 for treatment of moderate-to-severe AD (Beck et al.
2014). Several other types of targeted therapeutics are either FDA approved or under investigation;
the latter group includes oral and topical forms of Janus kinase (JAK) inhibitors. This chapter will
discuss clinical and histopathologic aspects of AD, and mechanisms of AD immunopathogenesis,
including the role of JAK/STAT signaling in AD, and conclude with a review of available data
regarding use of JAK inhibitors in human and canine AD.

5.2 Clinical Characteristics of AD


Based on estimates of up to 17% of children in the United States developing atopic dermatitis
(AD) and 2009 U.S. census data, more than 13 million children in the United States have AD
(Laughter et al. 2000). The worldwide prevalence of AD in adults is 1–3%, equating to millions of
additional cases (Katsarou and Armenaka 2011). There are two different phases of AD lesions,
acute and chronic (Rajka 1986). Acute lesions are red (erythematous), scaly, minimally elevated
plaques with poorly defined borders. In contrast to acute lesions, chronic lesions are more well
defined, often less scaly, and are more elevated appearing with accentuated skin lines (i.e.
lichenified). Chronic lesions develop from acute lesions, in part as a result of repetitive rubbing
and scratching of the skin (Rajka 1986). The areas of the body affected by AD (i.e. the
distribution) change with age. In infancy, the flexor aspects of the extremities and face are more
prominently involved. In contrast, the involvement of the extensor aspects of the extremities is

69
70 JAK-STAT Signaling in Diseases

more common later in childhood and adulthood. As implied above, the extent of skin affected
by AD patients can be limited to specific body regions; however, disease can be much more
extensive involving the majority of a patient’s skin. In addition to AD lesions themselves, AD
patients often have other associated clinical findings including dry skin (xerosis), ichthyosis
vulgaris (fish-like scaling of the skin caused by autosomal semidominant mutations in the
Filaggrin gene) and keratosis pilaris as well as comorbid allergic diseases such as asthma, allergic
rhinitis and food allergies (Rajka 1986; Smith et al. 2006).
As noted in the introduction, severe itch (pruritus), poor sleep and secondary skin infections are
major sources of morbidity in AD. Regardless of the extent of skin involvement, itch is one of the
defining features of AD (Rajka 1986). Patients often respond to itch by scratching the skin, which can
have several effects including: minor bleeding, inoculation of injured skin with microbial pathogens,
promotion of acute to chronic lesion conversion, and development of scarring and/or skin dyspig­
mentation. In addition, itching and scratching are important contributors to sleep disruption in AD
(Stores et al. 1998). With respect to secondary skin infections, Staphylococcus aureus, which also
frequently colonizes skin in AD patients without causing disease, is probably the most common
pathogen (Ong and Leung 2016; Totte et al. 2016). Typically, S. aureus infections manifest as
impetigo, a superficial skin infection that may resolve even without antibiotic therapy if a patient’s AD
is treated with anti-inflammatory agents (Travers et al. 2012). In contrast, herpes simplex virus
infections of AD skin lesions (eczema herpeticum) benefits from prompt treatment with systemic
antiviral therapies, particularly in the case of those that put the patient at risk of developing eye
infections, which can result in permanent vision loss (Young et al. 2010). As a result of the
aforementioned morbidities, the quality of life in AD can be dramatically impaired. For example,
children with generalized AD reported greater impairment in quality of life than patients with
asthma, diabetes, and epilepsy (Beattie and Lewis-Jones 2006). Healthcare costs for AD are also
significant (Mancini et al. 2008). In light of the above and the historical lack of widely effective
treatments for moderate-to-severe disease, patients and the medical community recognize the need to
develop more effective therapeutic agents for AD.

5.3 Histologic Aspects of AD


This section begins with a review of the histologic architecture of healthy skin and the histopathologic
changes seen in AD. These findings are summarized in Figure 5.1. Although the architecture varies
between body regions, the skin can generally be thought of as being composed of three layers, the upper
layer (epidermis), a middle layer (dermis) and the underlying subcutaneous fat. The outermost layer of
the skin, the epidermis, is also composed of multiple layers as detailed below. While this simplistic three-
layer concept of the skin will be discussed below for understanding AD pathogenesis, it is worth noting
that the skin is a highly specialized and complex organ system composed of many resident cell types and
structures including nerves, blood vessels, hair follicles, eccrine glands, apocrine glands, and others. The
role of all of these structures and cell types in AD is beyond the scope of this discussion.
The lowest (basal) layer of the epidermis is composed of stem cells and undifferentiated keratinocytes.
These undifferentiated keratinocytes leave the basal layer and transit upward through the epidermis,
undergoing progressive and eventually terminal differentiation, at which point their nucleus is lost and
they become protein-rich devitalized structures called corneocytes that are embedded in a lipid-rich
matrix. This composite of corneocytes and lipid matrix is the uppermost layer of skin called the stratum
corneum, a key barrier structure that prevents harmful irritants, allergens, and pathogens from entering
the skin and promotes water retention in the skin (Matsui and Amagai 2015).
Histopathologic changes in AD include increased thickness of the stratum corneum (ortho- and
parakeratosis), increased epidermal thickness (acanthosis) due to epidermal hyperplasia, spongio­
sis due to intercellular edema in the epidermis, and lymphocytic and histiocytic inflammatory cell
infiltration into the dermis and epidermis. In comparing acute and chronic AD lesions, acanthosis
and inflammatory infiltrates are more prominent and spongiosis is often less prominent in chronic
Role of JAK-STAT Signaling in Atopic Dermatitis 71

FIGURE 5.1 Basic histologic features of healthy human skin and atopic dermatitis lesional skin. The epidermis is
predominately composed of keratinocytes. Other cells in the epidermis include melanocytes and Langerhans cells (not
shown). The four layers of the epidermis include the stratum basale (SB), stratum spinosum (SS), stratum granulosum
(SG), and the stratum corneum (SC). The SB is the proliferative and least differentiated layer. Keratinocytes in the SG
have keratohyalin granules (purple dots), which contain filaggrin. The SC is the devitalized terminally differentiated
layer of the epidermis and is composed of corneocytes embedded in a lipid matrix. The SC and SG are major
contributors to epidermal barrier function. Spongiosis, which is caused by intercellular edema leading to keratinocyte
separation, is a common histologic feature of atopic dermatitis; however, spongiosis is not always observed in chronic
atopic dermatitis lesions. A perivascular lymphohistiocytic infiltrate is also seen in atopic dermatitis lesional skin; the
density and composition of this infiltrate varies between acute and chronic atopic dermatitis lesional skin and different
atopic dermatitis endotypes. Other histologic features that can be seen in atopic dermatitis lesional skin include
parakeratosis (thickened SC composed of corneocytes with retained nuclei), hypergranulosis (increased numbers of
keratohyalin granules in the SG), and lymphocyte exocytosis (lymphocytes in epidermis). Of note, the histologic
features of atopic dermatitis are not distinguishable from many other types of eczematous dermatitis.

lesions. Immunohistochemical and other analyses demonstrate that T cells are abundant in the
epidermal and dermal inflammatory cell infiltrates (Zachary et al. 1985).

5.4 Pathogenesis of AD
The pathogenesis of AD is mediated by a complex interplay between skin barrier dysfunction and
immune system dysregulation (Werfel et al. 2016). Atopic dermatitis is a complex genetic disease;
Filaggrin gene mutations are perhaps the best-characterized genetic defect in AD (Palmer et al.
2006; Weidinger et al. 2006). There is also a great deal of genetic, clinical, histologic, and
immunologic variation between subjects with AD. For example, most AD patients do not
harbor Filaggrin gene mutations, and there is variation in the type and frequency of Filaggrin
mutations between different populations (Akiyama 2010). The presence (extrinsic AD) or absence
(intrinsic AD) of elevated total serum IgE levels provide another was to subdivide AD (Novak
and Bieber 2003). There are also differences in clinical, histologic and immunologic characteristics
of AD in certain Asian populations who can manifest with a more psoriasiform phenotype
compared to classical AD features commonly seen in those of European descent (Noda et al.
2015). Taken together, these observations highlight the heterogeneity of AD and support the
72 JAK-STAT Signaling in Diseases

concept that there are subtypes (endotypes) of disease (Werfel et al. 2016). However, there
are characteristic clinical, histologic, cellular, and molecular aspects of disease that help
explain AD pathogenesis. As this chapter is focused on JAK-STAT signaling and JAK inhibition
as a potential treatment of AD, much of what follows in this section examines the immunologic
aspects of AD pathogenesis, some of which are summarized in Figure 5.2.

5.4.1 Epidermal Barrier Dysfunction


Development of AD involves two key and interrelated processes, epidermal barrier dysfunction
and T helper type 2 (Th2) cell-polarized immune responses (Bieber 2008; Sehra et al. 2008b; De
Benedetto et al. 2009). The stratum corneum is the outermost part of the epidermal barrier and is
composed of devitalized keratinocytes (i.e. corneocytes) embedded in a lipid matrix (Elias and
Schmuth 2009; Kypriotou et al. 2012). Cornified envelope proteins including filaggrin, loricrin and

FIGURE 5.2 Immunopathogenic mechanisms of atopic dermatitis that involve JAK-STAT signaling. In AD, a variety
of cytokines are produced by keratinocytes (TSLP, IL-33), Th2 cells (IL-4, IL-5, IL-13 and IL-31), and Th22 cells
(IL-22). With the exception of IL-33, these cytokines activate receptors (green) that employ JAK-STAT pathways. Top
panel: the upper layer of the skin (epidermis) is composed of keratinocytes that undergo progressive differentiation. As
keratinocytes differentiate and progress upward through the epidermis, they acquire the ability to produce the
epidermal barrier protein, filaggrin (FLG), which is stored in intracellular granules (purple circles). Keratinocytes
then lose their nucleus and become devitalized corneocytes embedded in a lipid matrix; this layer is called the stratum
corneum (SC) and forms a key part of the skin barrier. Stimulation of keratinocytes with IL-4 reduces keratinocyte
maturation, filaggrin expression, and proper stratum corneum formation. In contrast, IL-22 stimulation of keratino­
cytes promotes epidermal hyperplasia. Middle panel: Stimulation of Th2 cells with IL-4, IL-33, and TSLP promotes
Th2 cell differentiation and/or cytokine production. Bottom panel: Stimulation of B cells with IL-4 promotes IgE class
switching and production. Peripheral eosinophilia is stimulated by IL-5 and IL-33. Cytokines that act directly on
neurons to stimulate itch include IL-4, IL-31, and TSLP.
Role of JAK-STAT Signaling in Atopic Dermatitis 73

involucrin are key structural and functional proteins in corneocytes and are critical for proper
epidermal barrier function (Elias and Schmuth 2009; Kypriotou et al. 2012). These proteins are
encoded within the epidermal differentiation complex (EDC) gene cluster on human chromosome
1q21 (Elias and Schmuth 2009; Kypriotou et al. 2012). Barrier dysfunction in AD has been
classically attributed to inherent defects (i.e. genetic) in the keratinocyte; the most well-established
example of such a defect is the presence of mutations in the Filaggrin gene (Palmer et al. 2006;
Weidinger et al. 2006).
Filaggrin is a cornified envelope protein that normally aggregates keratin filaments in
corneocytes facilitating the ordered formation of a ‘bricks and mortar’ type structure of the
stratum corneum and an effective barrier that helps impede water loss from the skin and entry
of certain microbes and environmental agents into the skin (Elias and Schmuth 2009; Scharsch­
midt et al. 2009; Kypriotou et al. 2012). Filaggrin mutations are detected in a significant
proportion of patients with AD and result in insufficient quantities of functional protein thereby
leading to defective bundling of keratin filaments and disarray in the organization of the
stratum corneum (Manabe et al. 1991; Brown and McLean 2012). Filaggrin breakdown
products are important components of the so-called natural moisturizing factor (NMF) that
acts as a humectant helping to retain water in the skin. In patients with Filaggrin gene
mutations and AD, decreased levels of filaggrin and as a result decreased levels of NMF, are
thought to contribute to skin dryness (Rawlings and Harding 2004; Kezic et al. 2008). As noted
earlier, Filaggrin gene mutations also cause ichthyosis vulgaris, thus explaining the association
between this scaling skin condition and AD.
The preceding discussion on Filaggrin shows that intrinsic defects in keratinocytes can and
often do occur in AD, resulting in defective epidermal barrier function. These primary defects in
keratinocyte function leading to impaired skin barrier function can in turn promote cutaneous
inflammation in AD (Kim and Leung 2018). Interestingly, the majority of patients with AD do
not harbor Filaggrin gene mutations (Margolis et al. 2012). This suggests the impaired keratino­
cyte differentiation and function in AD do not necessarily require intrinsic defects in keratino­
cytes. One explanation for this observation is that in AD, the Th2 cytokine IL-4 signals through
its cognate receptor on keratinocytes leading to impaired keratinocyte differentiation and epider­
mal barrier dysfunction (Howell et al. 2007; Sehra et al. 2010). This suggests that inflammation is
sufficient to induce keratinocyte and skin barrier dysfunction in AD. Finally, environmental
insults can induce epidermal barrier disruption and resultant production of keratinocyte-derived
cytokines (e.g. TSLP, IL-33), which can in turn promote Th2-polarized inflammation in AD
(Oyoshi et al. 2010). Thus, extrinsic and intrinsic pathways can promote keratinocyte/epidermal
barrier dysfunction and immune dysregulation. Furthermore, there is a bidirectional positive
feedback loop between keratinocyte/epidermal barrier dysfunction and Th2 cytokine-mediated
immune dysregulation in AD pathogenesis. These data suggest a central role for IL-4-mediated
signaling in AD and suggest JAK inhibition could counteract the ability of IL-4 to impair
keratinocyte function in this disease.

5.4.2 T Cell Subsets and Cytokines


Histologic analyses of nonlesional skin from patients with AD reveal sparse perivascular T cell-
rich inflammatory infiltrates, while acute lesional skin reveals denser T cell-rich inflammatory
infiltrates of CD4+ and less abundant CD8+ T cells (Zachary et al. 1985). In acute lesions, Th2
cells and Th2 cytokines such as IL-4, IL-5 and IL-13 are predominant. In chronic lesions, Th2
cells and cytokines persist, but the T cell infiltrate becomes more diverse to include Th1 cells
(producing IFNγ) and other subsets (Hamid et al. 1994, 1996; Leung and Soter 2001). Clinical
investigations using atopy patch testing to mimic AD lesion formation revealed similar patterns of
T helper cell infiltrates with an early influx of Th2 cells and a later accumulation of Th1 cells
(Grewe et al. 1998). Th2 cytokines contribute to AD pathogenesis through multiple mechanisms
including promotion of Th2 cell differentiation, class switching and IgE production by B cells,
74 JAK-STAT Signaling in Diseases

eosinophil recruitment into the skin and decreased EDC gene expression by keratinocytes (Howell
et al. 2007; Kim et al. 2008; Sehra et al. 2010; Carmi-Levy et al. 2011). Th2 cells can also produce
IL-31, which can activate sensory neurons causing itch (Bilsborough et al. 2006; Sonkoly et al.
2006; Takaoka et al. 2006). As IL-4, IL-5, IL-13, and IL-31 receptors activate JAK/STAT
signaling, inhibition of JAK activity at these receptors could theoretically counteract the effects
of these cytokines thereby attenuating Th2 cell differentiation and cytokine production, IgE
production, peripheral eosinophilia, and itch while enhancing keratinocyte differentiation in
subjects with AD.
Other T cell populations are also implicated in AD pathogenesis. Subsets of CD4+ and CD8+
T cells designated Th22 and Tc22 cells, respectively, are implicated in AD pathogenesis. Inter­
leukin (IL)-22 produced by these cells is thought to promote the epidermal hyperplasia seen
in AD and other cutaneous inflammatory diseases like psoriasis (Nograles et al. 2009). An IL-9
producing T helper population (Th9) has also been identified in lesional skin of patients with AD,
though the role of these cells in disease is still under investigation (Schlapbach et al. 2014).
Regulatory T cells may also mediate AD pathogenesis as an AD-like eczematous dermatitis
develops in patients with immunodysregulation polyendocrinopathy enteropathy X-linked
(IPEX) syndrome, a disease characterized by defective Treg development and function (Halabi-
Tawil et al. 2009). However, more work is required to understand if and how skin-derived
regulatory T cells are involved in AD pathogenesis. In conclusion, involvement of T cells in AD
is well established, but much work remains to understand the effector mechanisms by which these
cells interact with one another and keratinocytes to regulate pathogenesis.

5.4.3 Epithelial Cytokines


Epithelial cytokines linked to AD pathogenesis include thymic stromal lymphopoietin (TSLP) and
IL-33 (Schmitz et al. 2005; Saenz et al. 2008; Meephansan et al. 2012; Milovanovic et al. 2012;
Savinko et al. 2012; Kim et al. 2013). Production of TSLP and IL-33 by keratinocytes is increased
in lesional skin of AD (Soumelis et al. 2002; Briot et al. 2009). Epidermal barrier disruption, toll-
like receptor 2 activation by membrane fragments of S. aureus, and inflammatory cytokines can
promote TSLP production by keratinocytes (Angelova-Fischer et al. 2010; Vu et al. 2010; Ziegler
2012). Stimulation of dendritic cells with TSLP also polarizes naïve T cells differentiation towards
Th2 cells and promotes production of the chemokines CCL17 and CCL22, which can act as
chemoattractants for Th2 cells (Soumelis et al. 2002). Moreover, TSLP stimulation of Th2, ILC2,
basophils, and mast cells promotes Th2 cytokine production (Hammad and Lambrecht 2015).
Recently, TSLP was shown to directly stimulate sensory neurons in the skin leading to pruritus
(itch) in mice injected with recombinant TSLP (Wilson et al. 2013; Turner and Zhou 2014). These
findings indicate that TSLP can directly promote allergic inflammation (via actions on leukocytes)
and itch (via actions on neurons). Similar to TSLP, IL-33 can activate Th2 and other cell types to
promote allergic inflammation. While contribution of IL-33 to AD pathogenesis is less clear,
transgenic overexpression of IL-33 in basal layer keratinocytes is sufficient to cause an AD-like
phenotype in mice (Imai et al. 2013). Although IL-33 receptor signaling does not utilize JAK/
STAT pathways, TSLP does; thus TSLP-mediated effects on promoting allergic inflammation and
itch could be therapeutic targets for JAK inhibition.

5.4.4 JAK/STAT Signaling in Preclinical Models of AD


With the exceptions of IL-25 and IL-33, the cytokines described above that are involved in AD
pathogenesis utilize JAK/STAT signaling. As noted earlier, the Th2 cytokine IL-4 is an important
mediator of AD pathogenesis. Activation of the IL-4 receptor promotes STAT6 phosphorylation
and downstream inflammatory cascades. To study the impact of STAT6 function in T cells, a T
cell-specific transgenic mouse expressing a constitutively active mutant human STAT6 protein,
STAT6VT, was generated; subsequent analyses determined these (Stat6VT) mice develop allergic
Role of JAK-STAT Signaling in Atopic Dermatitis 75

skin inflammation referred to hereafter as an AD-like phenotype (Sehra et al. 2010). Scaly
erythematous papules and plaques that can develop on the head, neck, trunk, extremities, and
tail characterize the AD-like phenotype in these mice. Two phases of disease, early and late
lesions, mirror the clinical, histologic, and cytokine profiles of human AD (DaSilva-Arnold et al.
2018). Histologic features of AD-like lesions in these mice include acanthosis and spongiosis and
a moderately dense perivascular and interstitial lymphocyte- and eosinophil-rich infiltrate (Bruns
et al. 2003; Sehra et al. 2008a, 2010; Turner et al. 2014). Interleukin 4 is a key mediator of disease,
as the AD-like phenotype in Stat6VT mice does not develop on an IL-4 deficient background
(Sehra et al. 2008a). Notably, Stat6VT mice develop an AD-like phenotype despite having
‘normal’ keratinocytes, providing a model to study the function of ‘normal’ keratinocytes in the
context of inflammatory cascades initiated by Th2 cells. The fact that these mice develop
an AD-like skin phenotype in the absence of any other genetic lesions or environmental interven­
tions demonstrates the importance of JAK/STAT signaling in this mouse model of AD. In addition
to STAT6 signaling, endogenous STAT3 activity was also required for this disease phenotype
(Stritesky et al. 2011).
Mouse models of AD have also been used to study the effects of JAK inhibitors on allergic skin
inflammation. For example, Amano and co-workers studied the effects of topical JTE-052 (0.5%
[wt/vol] in acetone with 1% [vol/vol] DMSO) on the NC/Nga mouse model of AD. Compared to
vehicle-treated controls, JTE-052-treated NC/Nga mice exhibited less severe clinical scores and
reduced transepidermal water loss as well as increased filaggrin levels and natural moisturizing
factor (NMF) concentrations in the epidermis. Conversely, phosphorylation of STAT3 and
STAT6 was reduced in the epidermis of JTE-052-treated NC/Nga mice. This study also demon­
strated topical JTE-052 increased filaggrin levels and NMF concentrations in healthy human skin
xenografted onto athymic nude mice. In addition, JTE-052 treatment of human skin equivalents
and human keratinocytes promoted keratinocyte differentiation even in the presence of IL-4 and
IL-13 (Amano et al. 2015). Tanimoto and co-workers also studied JTE-052 in a mouse model
of AD that is induced by intradermal injection of ear skin with TSLP; systemic administration of
JET-052 also reduced AD severity (i.e. ear thickness) in a dose-dependent manner in this model
(Tanimoto et al. 2018).
In summary, there is a significant body of evidence that JAK/STAT signaling contributes
to AD-like skin inflammation. Studies with the Stat6VT mouse model demonstrate the
sufficiency of constitutive STAT6 activity to cause AD-like inflammation. Conversely, inhibi­
tion of Janus kinase activity with JTE-052 in mouse models of AD reduced phosphorylation
of STAT6 and STAT3, reduced disease severity, and improved several markers of skin barrier
function. Interestingly, JAK inhibition even enhanced skin barrier function in healthy skin.
This latter observation may reflect an ability of JTE-052 to inhibit the basal IL-4 stimulation
of keratinocytes that occurs in mouse epidermis under homeostatic conditions (Sehra et al.
2010). These preclinical studies provide mechanistic insights into how JAK inhibition could
impact AD pathogenesis.

5.5 Clinical Experience with JAK Inhibitors for AD


As discussed above, multiple cytokine signaling pathways and cell types implicated in AD
pathogenesis utilize JAK/STAT signaling. This coupled with studies of JAK/STAT signaling in
preclinical models of AD provide a strong rationale for the potential utility of JAK inhibition
for the treatment of AD. Currently, no JAK inhibitors are FDA-approved for treatment
of AD in humans. In contrast, the JAK inhibitor, oclacitinib, was FDA-approved in 2013 for
treatment of canine AD. The first half of this section will discuss data on use of JAK
inhibitors for treatment of AD in humans. The second half will discuss canine AD and use
of oclacitinib for treatment of this condition. Table 5.1 summarizes some of the data presented
in the text below.
76 JAK-STAT Signaling in Diseases

TABLE 5.1
Summary of JAK Inhibitors Under Investigation for Use in Atopic Dermatitis
Target
Drug Route Dosing population Study type Efficacy

Tofacitinib Oral 5 mg QD-BID Human Case SCORAD ↓66%; Itch ↓ 76.3%;


series Sleeplessness ↓ 100% (at 8–29 weeks)
Topical 2% oint Human IIa EASI ↓ 81.7% vs. 29.9% for placebo (at 4 weeks)
Baricitinib Oral 2–4 mg QD Human II EASI ↓ 64-65% vs. 46% for placebo (at 16 weeks)
SCORAD ↓ 41-47% vs. 21% for placebo
JTE-052 Topical 0.25, 0.5, 1, Human II Modified EASI ↓ 41.7–72.9% vs. 12.2% for
3% ointment BID placebo
and 62% for tacrolimus ointment (at 4 weeks)
Ruxolitinib Topical 1.5% ointment Human IIb Not currently available in peer-reviewed
BID publication
Upadacitinib Oral 7.5, 15, 30mg QD Human IIb Not currently available in peer-reviewed
publication
Oclacitinib Oral 0.4–0.6 mg/kg Dog RCT CADESI-02 ↓ 48.4 vs. 3.6% for placebo (at
BID →QD 4 weeks)
Itch ↓ 47.4 vs. 10.4% for placebo
Dog Open Quality of life ↑ 91%
label

5.5.1 JAK Inhibition for Treatment of AD in Humans


There are currently no JAK inhibitors approved for any dermatologic indications in humans (Shreberk-
Hassidim et al. 2017). Oral and/or topical formulations of tofacitinib, baricitinib, JTE-052,
ruxolitinib, and upadacitinib have been and/or are under investigation for treatment of AD. As
data for use of ruxolitinib and upadacitinib in AD have not been published, the section will focus
on studies of tofacitinib, baricitinib, and JTE-052. Prior to discussing the safety and efficacy
data related to these agents, the next paragraph will review the metrics used to evaluate AD in
clinical trials.
There are multiple scoring systems used to evaluate AD in clinical research. The two most
common disease scoring instruments are the eczema area and severity index (EASI) and scoring
atopic dermatitis (SCORAD) system. The EASI score calculation requires estimation of the
surface area affected by AD in a given body region (i.e. head/neck, trunk, upper limbs, lower
limbs) and the severity of several clinical parameters of AD (i.e. redness, induction, excoriation,
and lichenification) in each of those body regions (Hanifin et al. 2001). Improvement in the EASI
score is often written as EASI-x where ‘x’ represents a percent reduction in disease severity from
baseline. For example, EASI-50 means a patient or group of patients (or study subjects) had
a 50% or greater improvement in their EASI score compared to their first EASI score. If 65% of
patients achieved EASI-50, that means 65% of patients had a 50% or greater improvement from
their initial EASI score. Similar to the EASI, the SCORAD calculation includes a determination
of the body surface involved by AD but differs from the EASI in that the SCORAD only
evaluates severity (intensity) of a ‘representative’ area of AD and also factors in the symptoms of
itch and sleeplessness (Kunz et al. 1997). Other clinical scoring systems focus on specific
symptoms like itch and quality of life, but will not be discussed further in this section.
Tofacitinib is the first JAK inhibitor reported for treatment of AD (Levy et al. 2015). Levy et al.
reported off-label use of oral tofacitinib to treat six consecutive patients with recalcitrant moderate-
to-severe AD. This medication was used due to the recalcitrant nature of these cases of AD. Patients
in this study were treated with tofacitinib 5 mg by mouth daily (one patient) and twice daily (five
patients), and their disease was assessed with the SCORAD index. Patients were allowed to continue
Role of JAK-STAT Signaling in Atopic Dermatitis 77

use of topical corticosteroids and topical calcineurin inhibitors as needed; one patient also received
low-dose oral corticosteroids (prednisone 5 mg by mouth twice daily) at the beginning of the study.
The average SCORAD measurement decreased by 54.5% (SCORAD 36.5 to 16.5) at the initial
follow-up visit (4–14 weeks after starting tofacitinib) and by 66% at the second follow-up visit
(8–29 weeks after starting tofacitinib). In addition, the severity of itch (pruritus score) decreased by
69.9% and 76.3% at the first and second follow-up visits, respectively; sleeplessness was reduced by
71.2% and 100% at those same time points. No adverse events were reported in this cohort of six
patients. Limitations of the study included the small sample size and lack of a placebo control arm.
There are no randomized placebo-controlled (RCT) trials of oral tofacitinib for AD.
Topical formulations of tofacitinib have been studied in AD. Bissonnette and co-workers
reported results from a phase IIa RCT of 69 adult AD patients treated twice daily with tofacitinib
2% ointment or vehicle control for 4 weeks (Bissonnette et al. 2016). Patients had mild-to­
moderate AD. Disease severity was evaluated using EASI score. After 4 weeks of treatment,
EASI scores improved by 81.7% in the tofacitinib-treated group versus 29.9% improvement in the
vehicle-treated group. Significant reductions in itch were also reported in the tofacitinib treatment
group. Treatment emergent adverse events occurred in 31.4% of tofacitinib-treated subjects and
55.9% of vehicle-treated subjects. In the tofacitinib treatment group, gastroenteritis occurred in
one subject, a second developed bronchitis, and a third developed a furuncle; two cases of
nasopharyngitis and one case of upper respiratory tract infection developed in each treatment
group. No severe or serious adverse events or deaths occurred in either treatment group.
Baricitinib was studied in a 16-week phase II randomized double-blinded placebo controlled trial of
124 adult subjects from the United States and Japan with moderate-to-severe AD (Guttman-Yassky
et al. 2018). Prior to randomization, all patients were treated for 4 weeks with topical corticosteroid
(triamcinolone 0.1% cream). Subjects were then randomized to a placebo group or one of two
baricitinib treatment groups, but triamcinolone use was permitted throughout the study. The first
baricitinib treatment group received 2 mg by mouth once daily; the second group received 4 mg by
mouth once daily. All three groups were treated with placebo or baricitinib for 16 weeks. Response to
therapy was evaluated with EASI and SCORAD instruments. At week 16, the mean reduction in
EASI scores from baseline was 64% and 65% in groups that were treated with 2 and 4 mg baricitinib,
respectively, and 46% in the placebo-treated group. The mean reduction in SCORAD scores was 41%
and 47% in groups that were treated with 2 and 4 mg baricitinib, respectively, and 21% in the placebo-
treated group. Treatment emergent adverse events occurred in 46% and 71% in 2 and 4 mg baricitinib
groups, respectively, and 49% in the placebo group. During the treatment phase of the study, one
serious event was reported (i.e. a benign polyp of the large intestine was detected in one patient in the
4 mg baricitinib group). Headache and elevated blood creatine phosphokinase were reported in
3–13% of subjects in the baricitinib-treated groups but not in the placebo-treated group. There was no
difference in frequency of infections between the groups that were treated with placebo and baricitinib.
No severe events or deaths occurred in any group.
Topical therapy with the JAK inhibitor JTE-052 was also studied in a Phase II study of 327 Japanese
subjects (16–65 years of age at enrolment) with moderate-to-severe AD (Nakagawa et al. 2018). Study
subjects were randomly assigned to one of the following six study groups: 0.25%, 0.5%, 1% and 3% JTE­
052 ointment, vehicle ointment or tacrolimus 0.1% ointment twice daily for four weeks. Therapeutic
response was evaluated using a modified EASI score (excludes evaluation of disease on head/neck
region). At week 4, the mean reduction in the modified EASI score from baseline for the JTE-052
treatment groups was 41.7% (0.25% ointment), 57.1% (0.5% ointment), 54.9% (1% ointment), and
72.9% (3% ointment). These improvements were all significantly greater than the 12.2% mean reduction
in the modified EASI score seen in the vehicle-treated group. The mean reduction in the tacrolimus
treated group was 62%, similar to that seen with the higher concentrations of JTE-052. Itch was also
evaluated using the pruritus numerical rating scale (NRS). Based on the NRS, daytime and night-time
itch were both significantly improved at 1 week in the groups treated with 0.5%, 1%, and 3% JTE-052
compared to vehicle-treated groups, with improvements in pruritus beginning on day 1. The percentage
of total adverse events in the groups treated with JTE-052 was 16% compared to 19.2% and 43% in the
vehicle- and tacrolimus-treated groups, respectively. All adverse events were mild or moderate severity
78 JAK-STAT Signaling in Diseases

with no serious or severe events, including death, occurring during the study. The most common adverse
event noted in the JTE-052 treatment group was nasopharyngitis (3.4%).

5.5.2 JAK Inhibition for Canine AD


While no JAK inhibitors have been FDA approved for treatment of AD in humans, oclacitinib,
marketed as Apoquel, was FDA approved in 2013 for treatment of AD in dogs (canine AD).
Canine AD is a common dermatologic disease; one study demonstrated 8.7% of dogs were
reported to have AD (Lund et al. 1999). Canine AD shares similarities to human AD with
respect to some clinical, histologic, and microbiologic features and treatment modalities used
(Marsella and Girolomoni 2009; Cosgrove et al. 2013; Gedon and Mueller 2018). Environmental
and food allergens also appear to have a more prominent role in contributing to skin inflamma­
tion in canine AD (Marsella and Girolomoni 2009). Although the pathogenic mechanisms of
canine AD differ in some respects to human disease, Th2 cytokines are important drivers of
allergic skin inflammation in both human and canine AD (Gedon and Mueller 2018).
Cosgrove et al. performed a randomized double-blinded placebo-controlled trial of oral oclacitinib
for canine AD (Cosgrove et al. 2013). The study used 299 dogs (12 months of age or older) with AD
and was divided into placebo, 0.4 mg/kg and 0.6 mg/kg oclacitinib treatment arms. Dogs first
received placebo or oclacitinib by mouth twice daily for 14 days, followed by once daily dosing
thereafter for 98 additional days (total treatment duration 112 days). Response to therapy was
determined periodically using the canine AD extent and severity index (CADESI-02) determined by
the clinician and the dog owner’s assessment of their dog’s itching (owner pruritus visual analogue
scale). Disease severity and itch were both significantly reduced in oclacitinib-treated dogs compared
to placebo-treated controls. For example, a 48.4% reduction in CADESI-02 score was noted on days
14 and 28 in oclacitinib-treated dogs compared to 1.7% and 3.6% reductions in placebo-treated dogs
on days 14 and 28. With respect to itching, pruritus scores on days 1, 2, 7, 14 and 28 improved by
29.5–66.7% in oclacitinib-treated dogs compared to 3.9–10.4% improvement in the placebo group.
Data analysis after day 28 was limited by the fact that most (>86%) placebo-treated dogs experienced
worsening itch and/or AD and were therefore transitioned to an open-label arm in which they
received oclacitinib. Cosgrove et al. also performed a long-term study of oclacitinib therapy in
canine AD in which dogs were followed for up to 683 days of treatment with oclacitinib; in this
study, dog owners reported 91% of dogs exhibited an improved quality of life (Cosgrove et al. 2015).
In the 2013 study by Cosgrove et al., the majority of placebo-treated dogs were switched to oclacitinib
by day 16 due to worsening AD. This switch limited the ability of the investigators to compare adverse
events between oclacitinib and placebo-treated groups beyond day 16. Adverse events between days 0
and 16 in the placebo and oclacitinib groups included diarrhea, vomiting, anorexia and lethargy; these
events affected less than 5% of dogs with no clear difference between study groups and resolved in the
majority of dogs despite continuation of treatment. Of the 299 dogs used in this study, 283 received at
least one dose of oclacitinib. Adverse events affecting at least 5% of oclacitinib-treated dogs included
pyoderma (12%), nonspecified dermal nodules (12%), otitis, (9.9%), vomiting (9.2%) and diarrhea (6%).
Leukopenia due to neutropenia occurred in some dogs but the percentage of dogs affected was not
reported. Although some variance in hematologic parameters was seen, cell counts remained in the
normal range, as did serum chemistry values and urinalysis parameters. In a long-term investigation of
oclacitinib therapy for canine AD, Cosgrove et al. studied dogs for up to 683 days of treatment with
oclacitinib (Cosgrove et al. 2015). Adverse events and laboratory values in oclacitinib-treated dogs were
similar to those reported in the author’s earlier short-term study.

5.6 Conclusions
After decades of studying atopic dermatitis, some key aspects of pathogenesis are clear. Skin
inflammation, skin barrier disruption, and itch are key manifestations and targets for treatment
Role of JAK-STAT Signaling in Atopic Dermatitis 79

of AD. Cellular mediators of these manifestations include epithelial cells (keratinocytes), Th2 and
other immune cells and neurons. Each of these cell types utilize JAK/STAT signaling to respond
to one or more cytokines implicated in AD. In light of this, there is a strong rationale for the
realized and potential utility of JAK inhibitors for treatment of this disease. In canine AD,
systemic therapy with the FDA-approved JAK inhibitor, oclacitinib, has helped revolutionize
treatment of AD in dogs. As clinical studies progress, it will be interesting to see how JAK
inhibitors impact treatment of AD in humans.

ACKNOWLEDGMENTS
MJT was supported by awards from the Dermatology Foundation (Physician Scientist CDA), the
Veterans Administration (VA CDA2; IK2CX001019) and the Showalter Trust as well as Indiana
University School of Medicine. MJT also thanks Dr. Stephen Shideler for his kind support. RMS
was supported by the Indiana Cutaneous Biology Training Program NIH T32 fellowship award
(T32AR062495).

REFERENCES
Akiyama, M. 2010. FLG mutations in ichthyosis vulgaris and atopic eczema: spectrum of mutations and
population genetics. Br J Dermatol 162 (3):472–477.
Amano, W., S. Nakajima, H. Kunugi et al. 2015. The Janus kinase inhibitor JTE-052 improves skin
barrier function through suppressing signal transducer and activator of transcription 3 signaling.
J Allergy Clin Immunol 136 (3):667–677.e7.
Angelova-Fischer, I., I. M. Fernandez, M. H. Donnadieu et al. 2010. Injury to the stratum corneum
induces in vivo expression of human thymic stromal lymphopoietin in the epidermis. J Invest
Dermatol 130 (10):2505–2507.
Beattie, P. E., and M. S. Lewis-Jones. 2006. A comparative study of impairment of quality of life in
children with skin disease and children with other chronic childhood diseases. Br J Dermatol 155
(1):145–151.
Beck, L. A., D. Thaci, J. D. Hamilton et al. 2014. Dupilumab treatment in adults with moderate-to-severe
atopic dermatitis. N Engl J Med 371 (2):130–139.
Bieber, T. 2008. Atopic dermatitis. N Engl J Med 358 (14):1483–1494.
Bilsborough, J., D. Y. Leung, M. Maurer et al. 2006. IL-31 is associated with cutaneous lymphocyte
antigen-positive skin homing T cells in patients with atopic dermatitis. J Allergy Clin Immunol 117
(2):418–425.
Bissonnette, R., K. A. Papp, Y. Poulin et al. 2016. Topical tofacitinib for atopic dermatitis: a phase IIa
randomized trial. Br J Dermatol 175 (5):902–911.
Briot, A., C. Deraison, M. Lacroix et al. 2009. Kallikrein 5 induces atopic dermatitis-like lesions through
PAR2-mediated thymic stromal lymphopoietin expression in Netherton syndrome. J Exp Med 206
(5):1135–1147.
Brown, S. J., and W. H. McLean. 2012. One remarkable molecule: filaggrin. J Invest Dermatol 132
(3 Pt 2):751–762.
Bruns, H. A., U. Schindler, and M. H. Kaplan. 2003. Expression of a constitutively active Stat6 in vivo
alters lymphocyte homeostasis with distinct effects in T and B cells. J Immunol 170 (7):3478–3487.
Carmi-Levy, I., B. Homey, and V. Soumelis. 2011. A modular view of cytokine networks in atopic
dermatitis. Clin Rev Allergy Immunol 41 (3):245–253.
Cosgrove, S. B., D. M. Cleaver, V. L. King et al. 2015. Long-term compassionate use of oclacitinib in dogs with
atopic and allergic skin disease: safety, efficacy and quality of life. Vet Dermatol 26 (3):171–179. e35.
Cosgrove, S. B., J. A. Wren, D. M. Cleaver et al. 2013. Efficacy and safety of oclacitinib for the control of
pruritus and associated skin lesions in dogs with canine allergic dermatitis. Vet Dermatol 24
(5):479–e114.
DaSilva-Arnold, S. C., A. Thyagarajan, L. J. Seymour et al. 2018. Phenotyping acute and chronic atopic
dermatitis-like lesions in Stat6VT mice identifies a role for IL-33 in disease pathogenesis. Arch
Dermatol Res 310 (3):197–207.
80 JAK-STAT Signaling in Diseases

De Benedetto, A., R. Agnihothri, L. Y. McGirt, L. G. Bankova, and L. A. Beck. 2009. Atopic dermatitis:
a disease caused by innate immune defects? J Invest Dermatol 129 (1):14–30.
Elias, P. M., and M. Schmuth. 2009. Abnormal skin barrier in the etiopathogenesis of atopic dermatitis.
Curr Allergy Asthma Rep 9 (4):265–272.
Gedon, N. K. Y., and R. S. Mueller. 2018. Atopic dermatitis in cats and dogs: a difficult disease for
animals and owners. Clin Transl Allergy 8:41.
Grewe, M., C. A. Bruijnzeel-Koomen, E. Schopf et al. 1998. A role for Th1 and Th2 cells in the
immunopathogenesis of atopic dermatitis. Immunol Today 19 (8):359–361.
Guttman-Yassky, E., J. I. Silverberg, O. Nemoto et al. 2019. Baricitinib in adult patients with moderate-to­
severe atopic dermatitis: a phase 2 parallel, double-blinded, randomized placebo-controlled
multiple-dose study. J Am Acad Dermatol 80 (4):913–921.
Halabi-Tawil, M., F. M. Ruemmele, S. Fraitag et al. 2009. Cutaneous manifestations of immune
dysregulation, polyendocrinopathy, enteropathy, X-linked (IPEX) syndrome. Br J Dermatol 160
(3):645–651.
Hamid, Q., M. Boguniewicz, and D. Y. Leung. 1994. Differential in situ cytokine gene expression in acute
versus chronic atopic dermatitis. J Clin Invest 94 (2):870–876.
Hamid, Q., T. Naseer, E. M. Minshall, Y. L. Song, M. Boguniewicz, and D. Y. Leung. 1996. In vivo
expression of IL-12 and IL-13 in atopic dermatitis. J Allergy Clin Immunol 98 (1):225–231.
Hammad, H., and B. N. Lambrecht. 2015. Barrier Epithelial Cells and the Control of Type 2 Immunity.
Immunity 43 (1):29–40.
Hanifin, J. M., M. Thurston, M. Omoto, R. Cherill, S. J. Tofte, and M. Graeber. 2001. The eczema area
and severity index (EASI): assessment of reliability in atopic dermatitis. EASI Evaluator Group.
Exp Dermatol 10 (1):11–18.
Holm, E. A., H. C. Wulf, H. Stegmann, and G. B. Jemec. 2006. Life quality assessment among patients
with atopic eczema. Br J Dermatol 154 (4):719–725.
Howell, M. D., B. E. Kim, P. Gao et al. 2007. Cytokine modulation of atopic dermatitis filaggrin skin
expression. J Allergy Clin Immunol 120 (1):150–155.
Imai, Y., K. Yasuda, Y. Sakaguchi et al. 2013. Skin-specific expression of IL-33 activates group 2 innate
lymphoid cells and elicits atopic dermatitis-like inflammation in mice. Proc Natl Acad Sci U S A 110
(34):13921–13926.
Katsarou, A., and M. Armenaka. 2011. Atopic dermatitis in older patients: particular points. J Eur Acad
Dermatol Venereol 25 (1):12–18.
Kezic, S., P. M. Kemperman, E. S. Koster et al. 2008. Loss-of-function mutations in the filaggrin gene
lead to reduced level of natural moisturizing factor in the stratum corneum. J Invest Dermatol 128
(8):2117–2119.
Kim, B. E., D. Y. Leung, M. Boguniewicz, and M. D. Howell. 2008. Loricrin and involucrin expression is
down-regulated by Th2 cytokines through STAT-6. Clin Immunol 126 (3):332–337.
Kim, B. E., and D. Y. M. Leung. 2018. Significance of skin barrier dysfunction in atopic dermatitis.
Allergy Asthma Immunol Res 10 (3):207–215.
Kim, B. S., M. C. Siracusa, S. A. Saenz et al. 2013. TSLP elicits IL-33-independent innate lymphoid cell
responses to promote skin inflammation. Sci Transl Med 5 (170):170ra16.
Kunz, B., A. P. Oranje, L. Labreze, J. F. Stalder, J. Ring, and A. Taieb. 1997. Clinical validation and
guidelines for the SCORAD index: consensus report of the European Task Force on atopic
dermatitis. Dermatology 195 (1):10–19.
Kypriotou, M., M. Huber, and D. Hohl. 2012. The human epidermal differentiation complex: cornified
envelope precursors, S100 proteins and the ‘fused genes’ family. Exp Dermatol 21 (9):643–649.
Laughter, D., J. A. Istvan, S. J. Tofte, and J. M. Hanifin. 2000. The prevalence of atopic dermatitis in
Oregon schoolchildren. J Am Acad Dermatol 43 (4):649–655.
Leung, D. Y., and N. A. Soter. 2001. Cellular and immunologic mechanisms in atopic dermatitis. J Am
Acad Dermatol 44 (1 Suppl):S1–S12.
Levy, L. L., J. Urban, and B. A. King. 2015. Treatment of recalcitrant atopic dermatitis with the oral
Janus kinase inhibitor tofacitinib citrate. J Am Acad Dermatol 73 (3):395–399.
Lund, E. M., P. J. Armstrong, C. A. Kirk, L. M. Kolar, and J. S. Klausner. 1999. Health status and
population characteristics of dogs and cats examined at private veterinary practices in the United
States. J Am Vet Med Assoc 214 (9):1336–1341.
Role of JAK-STAT Signaling in Atopic Dermatitis 81

Manabe, M., M. Sanchez, T. T. Sun, and B. A. Dale. 1991. Interaction of filaggrin with keratin filaments
during advanced stages of normal human epidermal differentiation and in ichthyosis vulgaris.
Differentiation 48 (1):43–50.
Mancini, A. J., K. Kaulback, and S. L. Chamlin. 2008. The socioeconomic impact of atopic dermatitis in
the United States: a systematic review. Pediatr Dermatol 25 (1):1–6.
Margolis, D. J., A. J. Apter, J. Gupta et al. 2012. The persistence of atopic dermatitis and filaggrin (FLG)
mutations in a US longitudinal cohort. J Allergy Clin Immunol 130 (4):912–917.
Marsella, R., and G. Girolomoni. 2009. Canine models of atopic dermatitis: a useful tool with untapped
potential. J Invest Dermatol 129 (10):2351–2357.
Matsui, T., and M. Amagai. 2015. Dissecting the formation, structure and barrier function of the stratum
corneum. Int Immunol 27 (6):269–280.
Meephansan, J., H. Tsuda, M. Komine, S. Tominaga, and M. Ohtsuki. 2012. Regulation of IL-33
expression by IFN-gamma and tumor necrosis factor-alpha in normal human epidermal
keratinocytes. J Invest Dermatol 132 (11):2593–2600.
Milovanovic, M., V. Volarevic, G. Radosavljevic et al. 2012. IL-33/ST2 axis in inflammation and
immunopathology. Immunol Res 52 (1–2):89–99.
Nakagawa, H., O. Nemoto, A. Igarashi, and T. Nagata. 2018. Efficacy and safety of topical JTE-052,
a Janus kinase inhibitor, in Japanese adult patients with moderate-to-severe atopic dermatitis: a phase
II, multicentre, randomized, vehicle-controlled clinical study. Br J Dermatol 178 (2):424–432.
Noda, S., M. Suarez-Farinas, B. Ungar et al. 2015. The Asian atopic dermatitis phenotype combines
features of atopic dermatitis and psoriasis with increased TH17 polarization. J Allergy Clin
Immunol 136 (5):1254–1264.
Nograles, K. E., L. C. Zaba, A. Shemer et al. 2009. IL-22-producing “T22” T cells account for
upregulated IL-22 in atopic dermatitis despite reduced IL-17-producing TH17 T cells. J Allergy
Clin Immunol 123 (6):1244–52.e2.
Novak, N., and T. Bieber. 2003. Allergic and nonallergic forms of atopic diseases. J Allergy Clin Immunol
112 (2):252–262.
Ong, P. Y., and D. Y. Leung. 2010. The infectious aspects of atopic dermatitis. Immunol Allergy Clin North
Am 30 (3):309–321.
Ong, P. Y., and D. Y. Leung. 2016. Bacterial and viral infections in atopic dermatitis: a comprehensive
review. Clin Rev Allergy Immunol 51 (3):329–337.
Oyoshi, M. K., R. P. Larson, S. F. Ziegler, and R. S. Geha. 2010. Mechanical injury polarizes skin
dendritic cells to elicit a T(H)2 response by inducing cutaneous thymic stromal lymphopoietin
expression. J Allergy Clin Immunol 126 (5):976–984. 984.e1–5.
Palmer, C. N., A. D. Irvine, A. Terron-Kwiatkowski et al. 2006. Common loss-of-function variants of the
epidermal barrier protein filaggrin are a major predisposing factor for atopic dermatitis. Nat Genet
38 (4):441–446.
Rajka, G. 1986. Natural history and clinical manifestations of atopic dermatitis. Clin Rev Allergy 4 (1):3–26.
Rawlings, A. V., and C. R. Harding. 2004. Moisturization and skin barrier function. Dermatol Ther 17
(Suppl 1):43–48.
Saenz, S. A., B. C. Taylor, and D. Artis. 2008. Welcome to the neighborhood: epithelial cell-derived
cytokines license innate and adaptive immune responses at mucosal sites. Immunol Rev 226:172–190.
Savinko, T., S. Matikainen, U. Saarialho-Kere et al. 2012. IL-33 and ST2 in atopic dermatitis: expression
profiles and modulation by triggering factors. J Invest Dermatol 132 (5):1392–1400.
Scharschmidt, T. C., M. Q. Man, Y. Hatano et al. 2009. Filaggrin deficiency confers a paracellular barrier
abnormality that reduces inflammatory thresholds to irritants and haptens. J Allergy Clin Immunol
124 (3):496–506. 506.e1–6.
Schlapbach, C., A. Gehad, C. Yang et al. 2014. Human TH9 cells are skin-tropic and have autocrine and
paracrine proinflammatory capacity. Sci Transl Med 6 (219):219ra8.
Schmitz, J., A. Owyang, E. Oldham et al. 2005. IL-33, an interleukin-1-like cytokine that signals via the
IL-1 receptor-related protein ST2 and induces T helper type 2-associated cytokines. Immunity
23 (5):479–490.
Sehra, S., H. A. Bruns, A. N. Ahyi et al. 2008a. IL-4 is a critical determinant in the generation of allergic
inflammation initiated by a constitutively active Stat6. J Immunol 180 (5):3551–3559.
82 JAK-STAT Signaling in Diseases

Sehra, S., F. M. Tuana, M. Holbreich et al. 2008b. Scratching the surface: towards understanding the
pathogenesis of atopic dermatitis. Crit Rev Immunol 28 (1):15–43.
Sehra, S., Y. Yao, M. D. Howell et al. 2010. IL-4 regulates skin homeostasis and the predisposition toward
allergic skin inflammation. J Immunol 184 (6):3186–3190.
Shreberk-Hassidim, R., Y. Ramot, and A. Zlotogorski. 2017. Janus kinase inhibitors in dermatology:
a systematic review. J Am Acad Dermatol 76 (4):745–753.e19.
Smith, F. J., A. D. Irvine, A. Terron-Kwiatkowski et al. 2006. Loss-of-function mutations in the gene
encoding filaggrin cause ichthyosis vulgaris. Nat Genet 38 (3):337–342.
Sonkoly, E., A. Muller, A. I. Lauerma et al. 2006. IL-31: a new link between T cells and pruritus in atopic
skin inflammation. J Allergy Clin Immunol 117 (2):411–417.
Soumelis, V., P. A. Reche, H. Kanzler et al. 2002. Human epithelial cells trigger dendritic cell mediated
allergic inflammation by producing TSLP. Nat Immunol 3 (7):673–680.
Stores, G., A. Burrows, and C. Crawford. 1998. Physiological sleep disturbance in children with atopic
dermatitis: a case control study. Pediatr Dermatol 15 (4):264–268.
Stritesky, G. L., R. Muthukrishnan, S. Sehra et al. 2011. The transcription factor STAT3 is required for
T helper 2 cell development. Immunity 34 (1):39–49.
Takaoka, A., I. Arai, M. Sugimoto et al. 2006. Involvement of IL-31 on scratching behavior in NC/Nga
mice with atopic-like dermatitis. Exp Dermatol 15 (3):161–167.
Tanimoto, A., Y. Shinozaki, Y. Yamamoto et al. 2018. A novel JAK inhibitor JTE-052 reduces skin
inflammation and ameliorates chronic dermatitis in rodent models: comparison with conventional
therapeutic agents. Exp Dermatol 27 (1):22–29.
Totte, J. E., W. T. van der Feltz, M. Hennekam, A. van Belkum, E. J. van Zuuren, and S. G. Pasmans.
2016. Prevalence and odds of Staphylococcus aureus carriage in atopic dermatitis: a systematic
review and meta-analysis. Br J Dermatol 175 (4):687–695.
Travers, J. B., A. Kozman, Y. Yao et al. 2012. Treatment outcomes of secondarily impetiginized pediatric
atopic dermatitis lesions and the role of oral antibiotics. Pediatr Dermatol 29 (3):289–296.
Turner, M. J., S. DaSilva-Arnold, N. Luo et al. 2014. STAT6-mediated keratitis and blepharitis: a novel
murine model of ocular atopic dermatitis. Invest Ophthalmol Vis Sci 55 (6):3803–3808.
Turner, M. J., and B. Zhou. 2014. A new itch to scratch for TSLP. Trends Immunol 35 (2):49–50.
Vu, A. T., T. Baba, X. Chen et al. 2010. Staphylococcus aureus membrane and diacylated lipopeptide
induce thymic stromal lymphopoietin in keratinocytes through the Toll-like receptor 2-Toll-like
receptor 6 pathway. J Allergy Clin Immunol 126 (5):985–993. 993.e1–3.
Weidinger, S., T. Illig, H. Baurecht et al. 2006. Loss-of-function variations within the filaggrin gene
predispose for atopic dermatitis with allergic sensitizations. J Allergy Clin Immunol 118 (1):214–219.
Werfel, T., J. P. Allam, T. Biedermann et al. 2016. Cellular and molecular immunologic mechanisms in
patients with atopic dermatitis. J Allergy Clin Immunol 138 (2):336–349.
Wilson, S. R., L. The, L. M. Batia et al. 2013. The epithelial cell-derived atopic dermatitis cytokine TSLP
activates neurons to induce itch. Cell 155 (2):285–295.
Young, R. C., D. O. Hodge, T. J. Liesegang, and K. H. Baratz. 2010. Incidence, recurrence, and outcomes
of herpes simplex virus eye disease in Olmsted County, Minnesota, 1976–2007: the effect of oral
antiviral prophylaxis. Arch Ophthalmol 128 (9):1178–1183.
Zachary, C. B., M. H. Allen, and D. M. MacDonald. 1985. In situ quantification of T-lymphocyte subsets
and Langerhans cells in the inflammatory infiltrate of atopic eczema. Br J Dermatol 112 (2):149–156.
Ziegler, S. F. 2012. Thymic stromal lymphopoietin and allergic disease. J Allergy Clin Immunol 130
(4):845–852.
6
JAK-STAT Signaling Pathway and Gliosis in
Neuroinflammatory Diseases

Han-Chung Lee
Laboratory Centre
Xiamen University Malaysia
Sepang, Malaysia

Kai-Leng Tan
Institute of Biomedical and Pharmaceutical Sciences
Guangdong University of Technology
Guangzhou, China

Pike-See Cheah
Department of Human Anatomy, Faculty of Medicine and Health Sciences
Universiti Putra Malaysia
Seri Kembangan, Malaysia

King-Hwa Ling
Department of Biomedical Sciences, Faculty of Medicine and Health Sciences
Universiti Putra Malaysia
Seri Kembangan, Malaysia
Department of Genetics
Harvard Medical School
Boston, Massachusetts United States

6.1 Introduction
It is commonly perceived that the brain composes of glia and neuron cells (GNR) in the ratio of
10:1. However, a recent review suggested that the GNR could be only about 1:1 (von Bartheld,
Bahney, and Herculano-Houzel 2016). In the human cerebral cortex, there are about 10–20 billion
neuronal cells, while the number of glial cells could be about 15–30 billion. There are three types
of glial cells in the human cerebral cortex with 45–75% of them are oligodendrocytes, 19–40% are
astrocytes, and 10% or less are microglia (von Bartheld, Bahney, and Herculano-Houzel 2016).
Each type of glial cell has a different role; for example, oligodendrocytes are the myelinating
cells that function to produce the insulating sheath of axons. On the other hand, microglia act as
immune cells in the central nervous system. In contrast, astrocyte has broader functions, which
include the maintenance of water and ion homeostasis, the blood–brain barrier, and modulation
of synaptic activity (Jakel and Dimou 2017; Dossi, Vasile, and Rouach 2018). However, reactive
microglia and astrocytes could have a detrimental effect on neuropathological conditions, for

83
84 JAK-STAT Signaling in Diseases

example, when they are stimulated by factors that are toxic to neurons such as β-amyloid and
α-synuclein, which are the critical molecule in the pathogenesis of Alzheimer’s disease (Hardy and
Higgins 1992) and Parkinson’s disease (Zhang et al. 2005), respectively.
Neuroinflammation is pathological inflammatory responses of the central nervous system
(CNS), which can lead to further damage/insult to the brain and spinal cord. The pathological
conditions in the brain include neurodegenerative diseases such as Alzheimer’s disease, Parkin­
son’s disease, and autoimmune diseases. These conditions induce both astrocytes and microglia,
which are the most significant and smallest neuroglial cells, respectively, to become reactive in
concert. Shreds of evidence now support that the reactive state of astrocytes and microglia are
deleterious and neurotoxic to the brain cells.
The Janus kinase–signal transducers and activators of transcription (JAK-STAT) signaling
pathway is an intracellular signaling pathway that involves the activation of two families of
proteins: The Janus kinases (JAK) and the signal transducer and activator of transcriptions
(STAT). JAK is a class of four cytoplasmic protein tyrosine kinases that includes JAK1, JAK2,
JAK3, and TYK2. The STAT family contains seven transcription factors: STAT1, STAT2,
STAT3, STAT4, STAT5A, STAT5B, and STAT6 (Kisseleva et al. 2002). JAK-STAT signaling
pathway is essential for the initiation of innate immunity, subsequently coordinating the adaptive
immune mechanism and ultimately constraining inflammatory responses (O’Shea and Plenge
2012). Specifically, the JAK-STAT signaling pathway is critical for T-cell proliferation, myeloid
and lymphoid cell differentiation, B cells, and macrophage activation. Under normal condition,
the protein expression level of JAK-STAT signaling pathway candidates such as JAK2, STAT1,
and STAT3 in the brain is low, which probably is due to the low cytokine secretion as the number
of immune cell present is small, and the glial cells are not in a reactive state (Na et al. 2007; Yan
et al. 2018).
Neurons, astrocytes, microglia, and immune cells are the essential sources for the cytokine
production in the brain (Galic, Riazi, and Pittman 2012). In general, cytokines are secreted in
response to infection or injury, and whether they are pro-inflammatory or anti-inflammatory
cytokines, the outcome will be different. Pro-inflammatory cytokines have been shown elevated
in many neurodegenerative diseases, which have a detrimental effect—such as damage to CNS
tissue—and are toxic to neurons and glial cells (Wang et al. 2015). Various cytokines have been
reported to activate the JAK-STAT signaling pathway, and different cytokines have the
propensity to activate a specific JAK and STAT. These include the interferons family such as
IFN-α, IFN-β, IFN-γ, IL-10, IL-19, IL-20, and IL-22; the gp130 family including IL-6, IL-11,
oncostatin M, leukaemia inhibitory factor, cardiotrophin-1, granulocyte colony-stimulating
factor, IL-12, leptin, ciliary neurotrophic factor, and cytokines under γ-chain family, such as
IL-2, IL-4, IL-7, IL-9, IL-15, and IL-21 (Rawlings, Rosler, and Harrison 2004; Schindler, Levy,
and Decker 2007).
Aberrant activation of the JAK-STAT pathway is becoming evident in contributing to neuroin­
flammatory diseases, with the involvement of activated microglia and astrocytes. Importantly, the
downregulation of JAK2–STAT3 signaling pathway was reported to cause memory impairment in
Alzheimer’s disease (Chiba et al. 2009), and potently promote the neurogenic-to-gliogenic shift in
Down syndrome brain (Lee et al. 2016). In this review, we aim to review the role of JAK-STAT
signaling pathway in neuroinflammation such as neurodegenerative disorders (e.g., Alzheimer’s
disease, Parkinson’s disease, Huntington’s disease), autoimmune disease (e.g., multiple sclerosis),
and developmental disorder due to chromosomal abnormalities (e.g., Down syndrome).

6.1.1 JAK-STAT and Reactive Astrogliosis and Microgliosis


Reactive gliosis is a pathological term referred to as the accumulation of astrocytes and microglia
at the injury site in the CNS (Streit, Mrak, and Griffin 2004). Advances in the neurobiology field
have pinpointed that glial cells, particularly astrocytes and microglia, are the essential players in
neuroinflammation (Heneka et al. 2015). Microglia are the resident phagocytes within the CNS
JAK-STAT Pathway and Neuroinflammation 85

(Heneka et al. 2015) and make up about 5–12% of the glial cell population (Gomez-Nicola and
Perry 2015). Generally, microglia exist in two distinct states, which are surveillance/resting state
versus activated state. Microglia in the resting state usually is elongated in shape with long
processes to facilitate its surveillance function. They are highly mobile, actively scouting their
niche by extending and retracting ramified processes while remaining stationary (Hristovska and
Pascual 2016). Upon detection of immunogen or pathogenic debris, microgliosis occurs and
initiates an immune response (Heneka et al. 2015).
When activated, they alter the cell morphology (Streit and Graeber 1993) and transform into
amoeboid shape with extensive cell processes toward the lesion site. It is the first-line defense
that protects the CNS after traumatic injury or infection (Gaudet et al. 2018). However, this
supposed beneficial mechanism when it is prolonged, becomes detrimental to neurons, with
or without the chronic persistence of the injury or immunogen (Cherry, Olschowka, and
Kerry O’Banion 2014). Activated microglia are further categorized into two different states:
M1 and M2 states. In neuroinflammation, the dynamic between M1 and M2 polarization is
crucial to determine the possible beneficial or detrimental outcomes (Cherry, Olschowka, and
Kerry O’Banion 2014). M1 microglia are responsive to harmful stimuli and secrete pro-
inflammatory cytokines such as TNF-α, IL-6, IL-1β, IFN-γ, and IL-23 that are known to
kick-start the inflammation cascades. In contrast, M2 microglia express inflammation inhibitory
cytokines, clear cellular debris through phagocytosis, and are capable of restoring homeostasis
(Hu et al. 2015).
JAK-STAT signaling pathway is the canonical pathway in immune and inflammatory responses.
It plays a vital role in microglial M1/M2 polarization (Hu et al. 2015). Extracellular signals such
as IFN-γ, IL-2, IL-6, and TNF-α trigger M1 polarization in microglia, while IL-4, IL-10, and
Galectin-1 stimulate M2 phenotype polarization. Members of the STAT family, including STAT1,
STAT3, and STAT6, are the intracellular switches in microglia M1/M2 phenotype interchange.
IFN-γ is found to induce M1 polarization via STAT1, while IL-13 and IL-4 stimulation polarize
microglia to M2 phenotype via STAT6 (Wang et al. 2015). Upon IFN-γ binds to its receptor,
JAK1/2 was activated, and STAT1 was phosphorylated, which subsequently trigger the expression
of inflammatory genes such as iNOS and IL-12. In contrast, JAK1/JAK3 or JAK1/Tyk2 was
activated by IL-4 or IL-13, respectively, which leads to phosphorylation of STAT6 and the release
of anti-inflammatory cytokine including IL-10 and IL-1RA (Tugal, Liao, and Jain 2013; Orihuela,
McPherson, and Harry 2016).
These microglial-secreted factors are acting through autocrine and paracrine manner to neurons
and astrocyte. The interplay between these cells can either further damage or repair the injury site.
Furthermore, cytokines (such as IFN-β and TNF-α) released from activated microglia are known
to stimulate astrocytes to secrete more inflammatory mediators including CCL2, CXCL1,
CXCL2, CXCL10, GM-CSF, and IL-6, as a secondary immune response (Pugazhenthi et al.
2013; Yan et al. 2018).
In response to pathological brain conditions, astrocytes exhibit cardinal features by becoming
hypertrophied and expressed a high level of GFAP, vimentin, and S100β (Zamanian et al. 2012).
This reactive state of astrocyte can be detrimental or beneficial to neuronal function, which
depends on the reactive state of the astrocyte, A1 or A2 type. Neuroinflammation-induced
astrocyte into A1 type led to downregulated genes such as Gpc6, Sparcl1, Mertk, and Megf10
that are critical for promoting synapse formation and phagocytic capacity (Liddelow et al. 2017).
A2 reactive astrocyte induced by ischemia upregulated many neurotrophic factors such as
CLCF1, LIF, IL6, and thrombospondins, which promote survival and growth of neurons, and
synapse repair (Zamanian et al. 2012; Liddelow et al. 2017).
Interestingly, activation of the JAK-STAT3 pathway has been reported as the central player in
the induction of reactive astrocytes, suggesting the pathway as a common mediator of astrocyte
reactivity was found highly conserved between disease states, species, and brain regions (Ben
Haim and Ceyzeriat 2015). Mounting evidence suggests that the JAK2–STAT3 pathway serves as
the critical regulator for astrocyte reactivity. See Figure 6.1 for a schematic illustration of
microglia and astrocyte activation in response to pathological brain conditions.
86

FIGURE 6.1 Microglia and astrocyte activation in response to pathological brain conditions. (A) Neuroinflammation process involving the activation of both resting microglia
and astrocytes into M1/M2 and A1/A2 states, respectively. (B) Multistep cellular activation of a JAK-STAT signaling pathway in microglia, astrocytes, or neurons upon the
activation by pro-inflammatory mediators. The activation of the JAK-STAT signaling pathway leads to the upregulation or downregulation of the transcription of targeted genes.
TSS: transcription start site.
JAK-STAT Signaling in Diseases
JAK-STAT Pathway and Neuroinflammation 87

6.1.2 JAK-STAT Signaling Pathway in Alzheimer’s Disease


Alzheimer’s disease (AD) is a progressive neurodegenerative disorder, which has been associated
with the extracellular deposition of neurotoxic β-amyloid (Aβ) plaques and intracellular neurofi­
brillary tangles (composed of hyperphosphorylated tau protein). There is ample evidence to show
that the deposition of Aβ neuritic plaques activates both the astrocytes and microglia to secrete
pro-inflammatory cytokines that are capable of inducing neuroinflammation (Lyman et al. 2014).
Pro-inflammatory cytokines such as IL-1β, IL-6, IFN-γ, TNF-α, and TNF-β have been reported
increased in human AD brain (Su, Bai, and Zhang 2016) and transgenic mouse models of AD
(Benzing et al. 1999; Apelt and Schliebs 2001). These cytokines exert an inflammatory effect
through activation of the JAK-STAT signaling pathway (Pugazhenthi et al. 2013; Rothaug,
Becker-Pauly, and Rose-John 2016), particularly the JAK2–STAT3 pathway. It was reported
that the activation of the JAK2–STAT3 pathway increased astrocyte reactivity, which in turn
increased β-amyloid deposition (Ceyzeriat et al. 2018). The mechanism is still largely unknown.
STAT3, however, has been associated with the β site APP cleaving enzyme 1 (BACE1), which is
responsible for Aβ generation (Wen et al. 2008). Besides, the upregulation of STAT3 promotes
inflammation genes expression such as iNOS and COX-2 and thus promotes neuroinflammation
(Lee et al. 2017).
Cytokines of the interleukins family activate the JAK-STAT signaling pathway. Upon
cytokine binding, the kinase JAK is activated and leads to STAT3 phosphorylation, which
dimerizes, translocated to the nucleus, and, subsequently, regulates the transcription of its
target genes. STAT3 activation in neurons is triggered in response to the increased release of
TNF-α from surrounding activated glial cells (Wan et al. 2010; Carret-Rebillat et al. 2015). It
is noteworthy that the upregulation of Tyk2 increases the phosphorylation of STAT3 and
contributes to neuronal apoptosis (Wan et al. 2010). The JAK-STAT signaling pathway was
found activated in microglial even before the first plaque deposition (Boza-Serrano et al.
2018). When amyloidogenesis continues and elevates Aβ plaque deposition, microglia
and astrocyte can be activated via toll-like receptor (TLR) and receptor for advanced
glycoxidation end products (RAGE)-dependent pathways. Subsequently, the activation of
caspases and signal-dependent transcription factors, such as NF-kB and activator protein 1
(AP-1) in microglia and astrocytes, leading to the release of pro-inflammatory cytokines
(Glass et al. 2010).
Furthermore, a pro-inflammatory cytokine, such as TNF-α, induced phosphorylation of STAT3,
which potently bound to 14-3-3 epsilon gene promoter and increased the expression that could lead
to more microglia activation (Xu et al. 2007; Tanabe et al. 2010; Zhang et al. 2012; Eufemi et al.
2015). Activated microglia also secrete several factors, IL-1α, TNFα, and complement component 1,
q subcomponent (C1q), which can induce more astrocytes into the reactive state (Figure 6.2)
(Liddelow et al. 2017). Indeed, activation of STAT3 promotes more microglia and astrocyte into
a reactive state. Reactive astrocytes are observed close to amyloid depositions in both patients
with AD (Probst, Ulrich and Heitz 1982) and mouse models of AD (Itagaki et al. 1989). Similar to
microglia, upon activated, reactive astrocytes reorganized arborization with increased number and
length of GFAP-positive processes (Wilhelmsson et al. 2006), polarize toward the site of amyloid
plaques aggregation (Nagele et al. 2003). It is unknown whether inhibition of STAT3 reduces β­
amyloid (Aβ) plaques aggregates. However, a study showed that suppression of STAT3 reduced the
expression of Aβ-induced gene transcription, such as iNOS, which markedly attenuates neuronal
cell death (Wan et al. 2010).

6.1.3 JAK-STAT Signaling Pathway in Parkinson’s Disease


Parkinson’s disease (PD) is an age-related and ranked as second most frequent chronic neurode­
generative disease worldwide (Darnell, Kerr, and Stark 1994; Troncoso-Escudero et al. 2018),
typically presented with decreased motor activity. Loss of dopaminergic neurons (DA) in the
substantia nigra pars compacta (SNc) and their projecting fibers in the striatum are observable
88 JAK-STAT Signaling in Diseases

FIGURE 6.2 Neuroinflammation in Alzheimer’s Disease.

Aβ neuritic plaques activate both the astrocytes and microglia to secrete pro-inflammatory cytokines such as IL-1β, IL-6,

IFN-γ, TNF-α, and TNF-β. The pro-inflammatory cytokines activate the JAK-STAT3 pathway, which, in turn, increases

the expression of 14-3-3 epsilon protein that leads to more microglia activation. Activated microglia secrete several

factors, IL-1α, TNFα, and complement component 1, q subcomponent (C1q), which induces more astrocytes into

a reactive state.

at the cellular level (Olanow and Tatton 1999). Reactive astrogliosis is observed upon neuroin­
flammation on various PD animal models, for example, familial PD associated-Nurr1 mutant
mice (Saijo et al. 2009), human alpha-synuclein (α-syn) stimulated mouse astrocytes (Fellner et al.
2013), and PD-related A53T mutant α-syn mice (Darnell, Kerr, and Stark 1994; Xing-Long et al.
2010). Overexpression of astrocytic A53T mutant α-syn caused reactive astrogliosis and microglial
activation accompanied by significant degeneration of DA and motor neurons in mice (Xing-Long
et al. 2010; Laurence et al. 2012). This neuroinflammation has long been considered a downstream
response to the death of dopaminergic neurons. However, mounting evidence suggests that astro­
cytes have an initiating role in PD pathophysiology.
The pathological hallmark of PD is the cytoplasmic inclusions of Lewy bodies that resulted from
misfolded and aggregated protein α-syn in presynaptic cells (Schulz-Schaeffer 2010). They are
proven to be toxic to DA (Petrucelli et al. 2002; Maries et al. 2003). Upon exposure to endogenous
stimuli, such as α-syn, resting microglia are activated and acquired the neurotoxic pro-inflammatory
M1 phenotype, which is deleterious to the DA neurons (Subramaniam and Federoff 2017). In PD
model, both wild type and mutant α-syn are found to potentiate microglia-driven neuroinflamma­
tion and further induce chronic progressive dopaminergic neurodegeneration (Gao et al. 2011,
Hoenen et al. 2016). JAK-STAT activation that promotes inflammatory M1 phenotype is proven
with the overexpression of STATs (STAT1, STAT2, STAT3) in α-syn stimulated microglia (Hoenen
et al. 2016, Qin et al. 2016) and α-syn overexpressed PD rat model (Qin et al. 2016). The pro-
inflammatory cascade activated in microglia and chronic PD progression is inhibited by the
inhibitors of inducible nitric oxide synthase (iNOS) and NADPH oxidase (Gao et al. 2011), which
JAK-STAT Pathway and Neuroinflammation 89

FIGURE 6.3 Summary of JAK-STAT expression in microglia M1/M2 phenotype polarization and released factors at

the presence of alpha-synuclein (α-syn) in PD models.

The presence of α-syn induced M1 phenotype in microglia via overexpression of STATs (STAT1, STAT2, and STAT3)

and release pro-inflammatory cytokines (IFN-γ, IL-6, IL-1β, TNF-α, and NO). Exposure of IL-4 induced M2

phenotype in microglia via the upregulation of STAT6 expression and facilitate phagocytosis clearance on α-syn.

are the two major free-radical synthesis enzymes in microglia. M2 microglia phenotype polarisation
is modulated by STAT6 activation via IL-4 stimulation (Park et al. 2016). This M2 polarised
microglia showed enhanced extracellular α-syn phagocytotic clearance that facilitate neuronal
survival (Park et al. 2016) (Figure 6.3).
Ample evidence indicates that the JAK-STAT pathway in neuroinflammation has a direct impact
on PD pathogenesis. Microglia release pro-inflammatory cytokines, including IL-1β, IL-6, TNF-α,
IFN-γ, and NO (Kawanokuchi et al. 2006; Nakagawa and Chiba 2015). IFN-γ and IL-6, two of the
most potent activators of the JAK-STAT pathway, are elevated in PD (Zhao et al. 2005; Chen et al.
2008; Gough et al. 2008; Yoda et al. 2010; Sherer 2011) and involved in the activation of the M1
phenotype. IFN-γ contributes to the degeneration of DA neurons through a mechanism involving
microglia. IFN-γ-deficient mice are protected against 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine
(MPTP) neurotoxicity with reduced loss of DA, striatal tyrosine hydroxylase, and dopamine
transporter fiber density (Zhao et al. 2005; Mount et al. 2007). In the presence of microglia,
exogenous IFN-γ ligand alone is enough to cause DA cell loss under neurotoxin treatment.
Whereas, IFN-γ receptor deficient-neurons are unable to resist cell death induced by a neurotoxin
(Zhao et al. 2005; Mount et al. 2007). These observations indicate the direct participation of IFN-γ­
induced microglial JAK-STAT on modulating DA loss in MPTP-induced PD models, despite the
existence of IFN-γ receptor on DA neurons. The consequence of the pro-inflammatory M1
microglia activation could be the key incident in neuroinflammation that leads to signature DA
loss observed in PD.
Downstream signaling pathways of M1 microglia could serve as important targets to inhibit the
pro-inflammatory damage. Human α-syn upregulates the expression of classical STAT-inducible
genes (iNOS, IL-6, TNF-α, MHC Class II, CIITA, and IRF-1) in primary mouse microglia (Qin
et al. 2016). AZD1480, a JAK1/JAK2 inhibitor, has been proven effective in inhibiting the JAK­
STAT signaling pathway and reducing STAT1 and STAT3 activation in α-syn-induced primary
mouse microglia (Nateghi Rostami et al. 2012; Qin et al. 2016). AZD1480 strongly inhibits the
subsequent overexpression of MHC Class II protein. In the same study, flow cytometry analysis
showed that AZD1480 suppressed microglial activation in the SNc of α-syn overexpressed PD rat
90 JAK-STAT Signaling in Diseases

model, compared to the baseline level in naïve rats. Most importantly, the stereological analysis of
tyrosine hydroxylase stained DA showed that AZD1480 attenuated the ~50% DA degeneration in
the SNc of AAV2-α-syn-transduced rat model after 3 months (Qin et al. 2016). RNAseq analysis
from the same sample also highlighted IFN-inducible genes, particularly IFNAR, STAT1, IFNB1,
IRF7, and TRIM24 that are regulated by the JAK-STAT signaling pathway in α-syn transduced
group. This proposes that repressing microglia activation via the JAK-STAT signaling pathway
might be beneficial to PD. It has been shown that in vivo LPS-induced microglia activation led to
PD-like symptoms, including the aggregation of α-syn and it was IL-1 dependent (Sakai et al.
1995; Tanaka et al. 2013). Collectively, the JAK-STAT signaling pathway in neuroinflammation is
a possible target to halt PD progression.

6.1.4 JAK-STAT Signaling Pathway in Huntington’s Disease


Huntington’s disease (HD) is a neurodegenerative disorder. It is characterized by progressive
motor dysfunction, psychiatric disorders, and cognitive deficits (Ross and Tabrizi 2011). The
disease is caused by a mutation in the huntingtin (HTT) gene (Träger et al. 2013; Crotti and Glass
2015). The CAG repeat expansion in the HTT gene encodes an expanded polyglutamine tract in
the huntingtin protein (HTT), resulting in a mutant protein. Mutant HTT protein with 40 or
more CAG repeats causes HD with full penetrance (MacDonald et al. 1993). While HD brain
quintessentially shows selective loss of neurons in the striatum and cortex, neuroinflammation has
been one of the leading clues to understand the pathogenesis of HD neurodegeneration. Notably,
the expression of mutant HTT increased microglial activation (Kraft et al. 2012). Immunostaining
of microglial marker, ionized calcium-binding adaptor molecule 1 (IBA1) in HD transgenic mice
striatum showed an increase of reactive microglia compared to wild-type mice (Simmons et al.
2007). Furthermore, mutant HTT expressing microglia promotes not only cell-autonomous
pro-inflammatory gene expression such as IL-6, but also triggers neighboring cells to secrete
pro-inflammatory cytokine (Crotti et al. 2014). The upregulation of several pro-inflammatory
cytokines, such as IL-6 and IL-8, was observed in the cerebrospinal fluid of HD patients (Bjorkqvist
et al. 2008). In addition, IL-6 upregulation was also reported in the cortex, cerebellum, and striatum
of HD patients (Silvestroni et al. 2009).
Increasing pro-inflammatory cytokine IL-6 potentially activates JAK-STAT3 signaling, which
in turn promotes astrocytic activation (Ben Haim and Ceyzeriat 2015; Luo and Zheng 2016).
The reactive astrocyte was observed in the striatum and cortex of patients with HD and HD­
transgenic mouse models. Besides, the increase of mutant HTT expression also contributes to
astrogliosis. A study showed that astrocytes were more hypertrophic and had increased STAT3
expression when the mutant HTT was delivered into the brain cells derived from mice and
monkeys (Ben Haim and Ceyzeriat 2015). Bioinformatics analysis identified STAT1 and STAT2
as potentially involved in the progression of HD. Protein–protein interaction network analysis
based on the differentially expressed genes (DEG) in the mutant huntingtin mouse cell line,
STHdh111/111, has identified 18 hub genes, including STAT1 and STAT2, due to the high
connectivity to the network; however, the study did not further investigate the role of these
genes in HD (Dong and Cong 2018). Besides, the analysis also found a novel regulatory
pathway (miR-124-STAT3-RNASE4) that could be a possible target for HD treatment (Dong
and Cong 2018). RNASE4 has been reported to have a neuroprotective effect, and it was
significantly upregulated in HD. MiRNA and transcription factor prediction showed that
RNASE4 is regulated by miR-124-STAT3 complex (Li et al. 2013; Dong and Cong 2018).
Taken together, the data showed that JAK-STAT signaling pathway is associated with the
pathogenesis of HD; however, to what extent that the pathway involves in HD neuroinflamma­
tion remains unknown.
Immunostaining analysis on post-mortem brain tissue from patients with HD showed that
many astrocytes in the brain were complement component 3 (C3)-positive, which is the specific
marker for A1 astrocytes (Liddelow et al. 2017). A1-reactive astrocytes lost the normal function
JAK-STAT Pathway and Neuroinflammation 91

and secreted many cytokines and chemokines, such as IL-6, chemokine CXCL-1, and IL-8 (Liu
et al. 2018). In consequence, this may attract circulating leukocytes to pass through the blood–
brain barrier and contribute to the chronic inflammatory progression in the brain (Liebner et al.
2018). On the other hand, the elevation of pSTAT5 level in HD monocytes has been reported
previously, which enhances the DNA-binding activity of NFκB leading to IL-6 production
(Kawashima et al. 2001; Träger et al. 2013). STAT5 does not directly activate the IL-6; it binds
to NFκB promoter and transactivates the promoter of IL-6, which has the NFκB-binding site
(Kawashima et al. 2001). Notably, the increased level of plasma IL-6 has been detected in HD
patients (Chang et al. 2015). Moreover, gene set enrichment analysis (GSEA) also found
differential expression of the JAK-STAT signaling pathway in HD monocyte (Miller et al. 2016).
Taken together, it is postulated that the elevation of circulating pro-inflammatory cytokine
released by HD monocyte infiltration into the brain and leading to chronic inflammation. In
brief, the existence of mutant HTT in the glial cells potentially induce neuroinflammation via
JAK-STAT signaling pathway in HD brains, both locally and systemically.

6.1.5 JAK-STAT Signaling Pathway in Multiple Sclerosis and Experimental


Autoimmune Encephalomyelitis
Multiple sclerosis (MS) is a demyelinating disease of the CNS with chronic neuroinflammation
and neurodegeneration (Constantinescu et al. 2011). MS exhibits four pathological features,
which are inflammation, demyelination, axonal loss, and gliosis. The pathological hallmark of
MS is demyelination, which is the destruction of the myelin sheath or the oligodendrocyte cell
body by the inflammatory process (Dendrou, Fugger, and Friese 2015). The etiology of MS
remains unknown, but it is widely accepted that neuroinflammation, specifically auto-reactive
lymphocytes, CD4+ T cells target CNS components and initiate the disease (Goverman 2009).
T helper type 1 (Th1) and T helper type 17 (Th17) lymphocytes produce IFN-γ and IL-17,
respectively (Constantinescu et al. 2011). In MS patients, T cells are peripherally primed by the
antigens, which possess molecular similarity with some CNS antigen (Sospedra and Martin 2005).
Later they differentiate into Th1 and Th17 cells that are capable of crossing the blood–brain
barrier (BBB) and causing a series of myelin-targeted autoimmune neuroinflammation. Other
immune cell types CD8+ T cells, B cells, and myeloid cells are also involved in promoting lesion
formation and neuronal damage (Yan et al. 2018).
Experimental autoimmune encephalomyelitis (EAE) is the animal model of MS. Auto-antigens,
such as myelin basic protein, proteolipid protein, and preferably myelin oligodendrocyte glyco­
protein (MOG), has been used to induce EAE in mice. Infiltrating macrophages and activated
microglia are the common populations found on the lesion site (Yan et al. 2018), and the EAE
model showed autoimmune response and MS-like clinical symptoms (Robinson et al. 2014). The
M1/M2 polarization of microglia and infiltrating macrophage is determined by the environmental
cytokine produced by Th1 cells and Th17 cells in EAE. The summary of how JAK-STAT
interplays between auto-reactive T cells and microglia polarization is shown in Figure 6.4. On
the infiltrating macrophage, IFN-γ secreted by Th1 cells promotes M1 polarization by activating
STAT1 in the JAK-STAT signaling pathway, while GM-CSF produced by Th17 cells promotes
M1 polarization via JAK2–STAT5 activation (Hamilton 2008). M1 macrophage polarization is
a disastrous event for MS and EAE progression. This has been proven by a study that showed
mice with GM-CSF receptor-deficient-CCR2+ monocytes are completely unresponsive from
EAE induction (Croxford et al. 2015). Subsequently, M1 macrophages produce IL-12 that
polarize T cells into Th1 cells via JAK2–TYK2 and STAT4 activation (Natarajan and Bright
2002; Krausgruber et al. 2011). Besides, IL-23 and IL-6 from M1 macrophages facilitate Th17
polarization by activating JAK1/2 and STAT3 in T cells (Samoilova et al. 1998; Cua et al.
2003). Thus, M1 macrophages are the kick starter of the vicious cycle in MS and EAE by
activating JAK-STAT signaling pathway within. From a different perspective, it is also shown
that macrophage and microglia are the most feasible candidates to halt the neuroinflammation
in MS or EAE. Studies showed that external stimuli, such as IL-4 (Francos-Quijorna et al.
92 JAK-STAT Signaling in Diseases

FIGURE 6.4 Summary of JAK-STAT expression in macrophage or microglia M1/M2 phenotype polarization and

released factors in MS and EAE models.

T lymphocytes transformed into autoreactive T helper type I (Th1) and Th17 lymphocytes in MS. Th1 and Th17

release IFN-γ and GM-CSF, respectively. These two factors induced M1 phenotype via STAT1 and JAK2–STAT5

activation. Pro-inflammatory cytokines, such as IL-12, IL6, and IL-23, further activate T cells into Th1 and Th17 cells.

IL-17 is Th17 autocrine factor. IL-4, IL-13, and IL-10 drive M2 phenotype and secrete IL-10 that are capable of

inactivating Th1 and Th17 auto-activation.

2016; Liu et al. 2016), IL-13 (Rutschman et al. 2001; Wynn, Chawla, and Pollard 2013), and
IL-10 (Lang et al. 2002; Gordon 2003), could drive them into M2 polarization. M2 macro­
phages and microglia exhibit strong anti-inflammatory reaction via an elevated level of IL-10
production (Moore et al. 2001; Lobo-Silva et al. 2016) reduced production of IL-23 (Correa
et al. 2011) and IL-6 (Sanchez-Ventura et al. 2019). Furthermore, M2 macrophage-secreted
IL-10 demonstrated a strong suppression on Th1 and Th17 phenotypes via STAT3 activation,
which then alleviate EAE symptoms (Qin et al. 2012; Liu et al. 2013; Li et al. 2018; Lotfi
et al. 2019). These showed that M2 macrophage and microglia are the pivots in relieving
clinical symptoms of MS and EAE. In brief, the JAK-STAT signaling pathway and the
responded cytokines or interleukins are utilized by both auto-reactive T cells and macro­
phages. It can be either deleterious or beneficial to the MS and EAE neuroinflammation, by
tweaking the M1/M2 phenotype polarization and the environmental cytokines via the JAK­
STAT signaling pathway.
Astrocytes are also responsive to environmental cytokines. They respond to IFN-β and secrete
suppressors of cytokine signaling (SOCS) via activation of the JAK-STAT signaling pathway to
negatively regulate immune cell infiltration, such as monocytes and CD4+ T cells, into the CNS (Qin
et al. 2008). IFN-β is a successful MS treatment as it reduces exacerbation rates and delays MS
progression (Weinstock-Guttman et al. 1995; Dhib-Jalbut 1997; Hohlfeld and Wekerle 2004). IFN­
β induced SOCS protein expression via STATs activation in primary murine astrocyte (Qin et al.
2008). The induction of SOCS-1 and SOCS-3 expression are dependent on STAT1α and STAT3,
respectively. In the same study, SOCS-1 and SOCS-3 silenced-astrocytes show greater chemoattrac­
tion to bone marrow-derived primary macrophages and CD4+ T cells under IFN-β stimulation
(Qin et al. 2008). This strengthens the importance of the JAK-STAT signaling pathway in
ameliorating neuroinflammation in MS. The depletion of IFN-γ receptors on astrocytes has
been shown to reduce the severity of EAE, which suggest that by ignoring IFN-γ detected in the
niche may lower the chance to trigger neuroinflammation (Ding et al. 2015). Similar to M1
macrophages, the secretion of IL-6 by astrocytes in response to endoplasmic reticulum stress
JAK-STAT Pathway and Neuroinflammation 93

(Meares et al. 2014) has been reported to be a determining factor of EAE neuroinflammation
(Quintana et al. 2009; Giralt et al. 2013). IL-6 deficient mouse is completely resistant to EAE
induced by MOG (Giralt et al. 2013). In the same study, the mouse with the only astrocyte
produced IL-6 demonstrated demyelination, leukocytes infiltration, and astrogliosis, restricted
within the cerebellar region, upon MOG stimulation. Like Th17 cells, it also produced GM-CSF
during the EAE process (Mayo et al. 2014) that will polarize macrophage into M1 phenotype via
JAK2–STAT5 (Hamilton 2008). Upregulated expression of lactosylceramide (LacCer) has been
reported in the CNS of EAE mice, and it regulates the astrocytes transcriptional program (Mayo
et al. 2014). In the same study, LacCer in astrocytes regulates the chemokine CCL2 and GM-CSF
production in a non-cell autonomous manner, which modulates the recruitment of activated
microglia and CNS-infiltrating monocytes. JAK-STAT in astrocytes is undeniably the central
control system in switching between beneficial and detrimental functions during neuroinflamma­
tion. Thus, the JAK-STAT signaling pathway might be the candidate to manipulate astrocyte to
counteract the progression of EAE and MS.

6.1.6 JAK-STAT Signaling Pathway in Down Syndrome


Very little is known about the neuroinflammation in the Down syndrome (DS) brain. However,
inflammatory receptor genes are found overexpressed in both human and mouse models for DS
(Ferrando-Miguel et al. 2003; Amano et al. 2004; Wilcock and Griffin 2013; Ling et al. 2014).
Notably, pro-inflammatory receptor genes located on chromosome 21, including IFNAR1, IFNAR2,
and IFNGR2 are triplicated in DS samples (Wilcock and Griffin 2013), and they can activate JAK­
STAT signaling pathway when interferons bind to the receptor. Once activated, the JAK-STAT
signaling pathway promotes astrocytes differentiation (Bonni 1997). Besides, the analysis of protein
expression level demonstrated that there was an approximately 2-fold increment of IFNAR2 expres­
sion in human fetal DS brain at 19–21 weeks of gestational age (Ferrando-Miguel et al. 2003). The
increased expression levels of Ifnar1, Ifnar2, and Il10rb genes were also seen in a DS mouse model,
Ts1Cje. These trisomic genes were overexpressed about 1.5-fold in Ts1Cje mouse whole brain (Amano
et al. 2004). In addition to Ifn receptor, the downstream target, STAT1, was also found upregulated in
the P84 Ts1Cje cerebellum and cerebral cortex (Ling et al. 2014). Furthermore, western blot analysis
on fibroblast cell lines from individuals with trisomy 21 confirmed the upregulation of IFN receptor
proteins (Sullivan et al. 2016).
The number of astrocytes is increased in human fetuses with DS. Immunohistochemical staining
on the frontal lobe of DS human fetuses showed increased GFAP positive cells as compared to
the age-matched controls (Zdaniuk et al. 2011). The finding was also consistent with previous
studies on both human fetuses and Ts65Dn mice brain (another mouse model for DS) (Contest­
abile et al. 2007; Guidi et al. 2008). Astrocytes in DS are not only more proliferative and
abundant, but they also display altered processes (Dossi, Vasile, and Rouach 2018).
In addition, pro-inflammatory cytokines, such as IL-6, TNF-α, and TGF-β, are well-known
activators of the JAK-STAT signaling pathway. These cytokines were at least 2-fold higher in
autopsied human DS brain tissues (<40 years old) when compared to age-matched controls. In
contrast, it showed at least a 3-fold increase in Down syndrome with AD (>40 years old)
(Wilcock et al. 2015). Besides, the level of IFN-γ was significantly increased in the whole brain
of the trisomy 16 mouse (Hallam et al. 2000). The presence of the extra copy of IFN receptor and
the elevation of IFN level have been postulated to sensitize the cells to interferon interaction and
lead to activation of the JAK-STAT signaling pathway (Ling et al. 2014; Lee et al. 2016).
Furthermore, it has been indicated that TNF-α and IFN-γ bind to receptors on microglia and
astrocytes (Benveniste and Benos 1995), which can activate both cell type and potently induces
JAK-STAT signaling pathway in microglia and astrocytes. Activation of glial cells includes
triggering of the JAK-STAT signaling pathway that could lead to neuroinflammation via the
release of nitric oxide (NO) (Figure 6.5) (Kim et al. 2002; Lively and Schlichter 2018). A previous
study reported that the gene expression level of inducible nitric oxide synthase (iNOS) that
stimulates NO generation was higher in DS astroglia than in control astroglia (Chen et al. 2014).
94 JAK-STAT Signaling in Diseases

FIGURE 6.5 Neuroinflammation in Down Syndrome.


When TNF-α and IFN-γ bind to receptors on microglia and astrocytes, it is potently to activate JAK-STAT signaling
pathway. Besides, an extra copy of IFN receptors on glial cells sensitizes the cells to interferon interaction, which
further triggers the JAK-STAT signaling pathway. In consequence, the expression level of iNOS increased in and
released NO, which leads to neuroinflammation.

In addition, increased of reactive oxygen species (ROS) production was observed in primary
cultures of hippocampal neurons and astrocytes derived from Ts1Cje, a mouse model for DS
(Shukkur et al. 2006). The expression of STAT1 and STAT3 were then highly induced in response
to ROS (Simon et al. 1998). In consequence, chronic expression of STAT1 and STAT3 stimulate
an inflammatory environment in the brain and lead to neuroinflammation. This could be through
upregulation of downstream transcriptional targets of STATs, such as Jmjd3, Ccl5, Ezr, Ifih1, Irf7,
Uba7, and Pim1, which are the genes that code for inflammatory proteins (Przanowski et al.
2014). JAK inhibitors or JAKinibs are currently a type of pharmacological approach that aids in
inhibiting the activity of one or more enzymes of the JAK family, thereby interfering with the
JAK-STAT signaling pathway. The application of JAKinibs, targeting astroglial and microglial
activation states in Down syndrome, offers excellent therapeutic potential and warrant further
investigation.

6.2 Concluding Remarks


The JAK-STAT signaling pathway in response to a wide range of cytokines is now known as
one of the critical factors to promote neuroinflammation in neurodegenerative disease. However,
it is conditional by whether it is pro-inflammatory cytokines or anti-inflammatory cytokines
that trigger the JAK-STAT signaling pathway. In neurodegenerative disease, M1-reactive
JAK-STAT Pathway and Neuroinflammation 95

microglial or A1-reactive astrocyte secretes pro-inflammatory cytokines that create an environ­


ment conducive to inflammation. Cytokines act on the cells that secrete them, or nearby cells,
and stimulate JAK-STAT signaling pathways. Activated JAK-STAT signaling leads to more
microglia change into the reactive state as well as increase more pro-inflammatory cytokines and
upregulated inflammatory gene expression. JAKinib is the inhibitor to one or more of the JAK
family, thereby suppressing JAK-STAT signaling pathway that could reduce the releasing pro-
inflammatory cytokines. However, inhibition of JAKs may potentially reduce anti-inflammatory
response as the JAK-STAT signaling pathway also triggers by anti-inflammatory cytokines. The
cytokines released in the microenvironment conditions the detrimental or beneficial effect;
therefore, modulation of the balance between pro-inflammatory and anti-inflammatory cyto­
kines, which would affect the function of the JAK-STAT signaling pathway, could be another
way to reduce inflammation.

ACKNOWLEDGEMENTS
This work was supported by Malaysian Ministry of Science, Technology and Innovation Science-
fund (Project ID: 02-01-04-SF2336) awarded to K.-H.L. and National Natural Science Founda­
tion of China (Project ID: 81850410549) awarded to K.-L.T.

REFERENCES
Amano, K., H. Sago, C. Uchikawa, et al. 2004. Dosage-dependent over-expression of genes in the
trisomic region of Ts1Cje mouse model for Down syndrome. Hum Mol Genet 13(13):1333–40.
Apelt, J., and R. Schliebs. 2001. Beta-amyloid-induced glial expression of both pro- and anti-inflammatory
cytokines in cerebral cortex of aged transgenic Tg2576 mice with Alzheimer plaque pathology. Brain
Res 894(1):21–30.
Ben Haim, L., and K. Ceyzeriat. 2015. The JAK–STAT3 pathway is a common inducer of astrocyte
reactivity in Alzheimer’s and Huntington’s diseases. J Neurosci 35(6):2817–29.
Benzing, W. C., J. R. Wujek, E. K. Ward, et al. 1999. Evidence for glial-mediated inflammation in aged
APP(SW) transgenic mice. Neurobiol Aging 20(6):581–89.
Benveniste, E. N., and D. J. Benos. 1995. TNF-alpha- and IFN-gamma-mediated signal transduction
pathways: effects on glial cell gene expression and function. Faseb J 9(15):1577–84.
Bjorkqvist, M., E. J. Wild, J. Thiele, et al. 2008. A novel pathogenic pathway of immune activation
detectable before clinical onset in Huntington’s disease. J Exp Med 205(8):1869–77.
Bonni, A. 1997. Regulation of gliogenesis in the central nervous system by the JAK–STAT signaling
pathway. Science 278(5337):477–83.
Boza-Serrano, A., Y. Yang, A. Paulus, and T. Deierborg. 2018. Innate immune alterations are elicited in
microglial cells before plaque deposition in the Alzheimer’s disease mouse model 5xFAD. Sci Rep 8
(1):1550.
Carret-Rebillat, A. S., C. Pace, S. Gourmaud, et al. 2015. Neuroinflammation and Abeta accumulation
linked to systemic inflammation are decreased by genetic PKR down-regulation. Sci Rep 5:8489.
Ceyzeriat, K., L. Ben Haim, A. Denizot, et al. 2018. Modulation of astrocyte reactivity improves

functional deficits in mouse models of Alzheimer’s disease. Acta Neuropathol Comm 6(1):104.

Chang, K. H., Y. R. Wu, Y. C. Chen, and C. M. Chen. 2015. Plasma inflammatory biomarkers for

Huntington’s disease patients and mouse model. Brain Behav Immun 44:121–27.
Chen, C., P. Jiang, H. Xue, et al. 2014. Role of astroglia in Down’s syndrome revealed by patient-derived
human-induced pluripotent stem cells. Nat Commun 5:4430.
Chen, H., E. J. O’Reilly, M. A. Schwarzschild, and A. Ascherio. 2008. Peripheral inflammatory
biomarkers and risk of Parkinson’s disease. Am J Epidemiol 167(1):90–95.
Cherry, J. D., J. A. Olschowka, and M. Kerry O’Banion. 2014. Neuroinflammation and M2 microglia: the
good, the bad, and the inflamed. J Neuroinflammation 11(1):98.
Chiba, T., M. Yamada, J. Sasabe, et al. 2009. Amyloid-beta causes memory impairment by disturbing the
JAK2/STAT3 axis in hippocampal neurons. Mol Psychiatry 14(2):206–22.
96 JAK-STAT Signaling in Diseases

Constantinescu, C. S., N. Farooqi, K. O’Brien, and B. Gran. 2011. Experimental autoimmune


encephalomyelitis (EAE) as a model for multiple sclerosis (MS). British J Pharmacol 164
(4):1079–106.
Contestabile, A., T. Fila, C. Ceccarelli, et al. 2007. Cell cycle alteration and decreased cell proliferation in
the hippocampal dentate gyrus and in the neocortical germinal matrix of fetuses with Down
syndrome and in Ts65Dn mice. Hippocampus 17(8):665–78.
Correa, F., M. Hernangomez-Herrero, L. Mestre, F. Loria, F. Docagne, and C. Guaza. 2011. The
endocannabinoid anandamide downregulates IL-23 and IL-12 subunits in a viral model of multiple
sclerosis: evidence for a cross-talk between IL-12p70/IL-23 axis and IL-10 in microglial cells. Brain
Behav Immun 25(4):736–49.
Crotti, A., C. Benner, B. E. Kerman, et al. 2014. Mutant Huntingtin promotes autonomous microglia
activation via myeloid lineage-determining factors. Nat Neurosci 17(4):513–21.
Crotti, A., and C. K. Glass. 2015. The choreography of neuroinflammation in Huntington’s disease.
Trends Immunol 36(6):364–73.
Croxford, A. L., M. Lanzinger, F. J. Hartmann, et al. 2015. The cytokine GM-CSF drives the
inflammatory signature of CCR2+ monocytes and licenses autoimmunity. Immunity 43(3):502–14.
Cua, D. J., J. Sherlock, Y. Chen, et al. 2003. Interleukin-23 rather than interleukin-12 is the critical
cytokine for autoimmune inflammation of the brain. Nature 421(6924):744–48.
Darnell, J. E., Jr., I. M. Kerr, and G. R. Stark. 1994. Jak-STAT pathways and transcriptional activation in
response to IFNs and other extracellular signaling proteins. Science 264(5164):1415–21.
Dendrou, C. A., L. Fugger, and M. A. Friese. 2015. Immunopathology of multiple sclerosis. Nat Rev
Immunol 15:545.
Dhib-Jalbut, S. 1997. Mechanisms of interferon beta action in multiple sclerosis. Mult Scler 3(6):397–401.
Ding, X., Y. Yan, X. Li, et al. 2015. Silencing IFN-gamma binding/signaling in astrocytes versus
microglia leads to opposite effects on central nervous system autoimmunity. J Immunol 194
(9):4251–64.
Dong, X., and S. Cong. 2018. Identification of differentially expressed genes and regulatory relationships
in Huntington’s disease by bioinformatics analysis. Mol Med Rep 17(3):4317–26.
Dossi, E., F. Vasile, and N. Rouach. 2018. Human astrocytes in the diseased brain. Brain Res Bull
136:139–56.
Eufemi, M., R. Cocchiola, D. Romaniello, et al. 2015. Acetylation and phosphorylation of STAT3 are
involved in the responsiveness of microglia to beta amyloid. Neurochem Int 81:48–56.
Fellner, L., R. Irschick, K. Schanda, et al. 2013. Toll-like receptor 4 is required for α-synuclein dependent
activation of microglia and astroglia. Glia 61(3):349–60.
Ferrando-Miguel, R., K. S. Shim, M. S. Cheon, M. Gimona, M. Furuse, and G. Lubec. 2003. Over-
expression of Interferon α/β receptor β chain in fetal Down syndrome brain. Neuroembryol Aging 2
(4):147–55.
Francos-Quijorna, I., J. Amo-Aparicio, A. Martinez-Muriana, and R. Lopez-Vales. 2016. IL-4 drives
microglia and macrophages toward a phenotype conducive for tissue repair and functional recovery
after spinal cord injury. Glia 64(12):2079–92.
Galic, M. A., K. Riazi, and Q. J. Pittman. 2012. Cytokines and brain excitability. Front Neuroendocrinol
33(1):116–25.
Gao, H. M., Zhang, F., Zhou, H., Kam, W., Wilson, B., Hong, J. S. 2011. Neuroinflammation and α­
synuclein dysfunction potentiate each other, driving chronic progression of neurodegeneration in a
mouse model of Parkinson’s disease. Environ Health Perspect 119(6):807–814.
Gaudet, A. D., L. K. Fonken, L. R. Watkins, R. J. Nelson, and P. G. Popovich. 2018. MicroRNAs: roles
in regulating neuroinflammation. The Neuroscientist 24(3):221–45.
Giralt, M., R. Ramos, A. Quintana, et al. 2013. Induction of atypical EAE mediated by transgenic
production of IL-6 in astrocytes in the absence of systemic IL-6. Glia 61(4):587–600.
Glass, C. K., K. Saijo, B. Winner, M. C. Marchetto, and F. H. Gage. 2010. Mechanisms underlying
inflammation in neurodegeneration. Cell 140(6):918–34.
Gomez-Nicola, D., and V. H. Perry. 2015. Microglial dynamics and role in the healthy and diseased brain:
a paradigm of functional plasticity. Neuroscientist 21(2):169–84.
Gordon, S. 2003. Alternative activation of macrophages. Nat Rev Immunol 3(1):23–35.
JAK-STAT Pathway and Neuroinflammation 97

Gough, D. J., D. E. Levy, R. W. Johnstone, and C. J. Clarke. 2008. IFNgamma signaling-does it mean
JAK–STAT? Cytokine Growth Factor Rev 19(5–6):383–94.
Goverman, J. 2009. Autoimmune T cell responses in the central nervous system. Nat Rev Immunol 9
(6):393–407.
Guidi, S., P. Bonasoni, C. Ceccarelli, et al. 2008. Neurogenesis impairment and increased cell death
reduce total neuron number in the hippocampal region of fetuses with Down syndrome. Brain
Pathol 18(2):180–97.
Hallam, D. M., N. L. Capps, A. L. Travelstead, G. J. Brewer, and L. E. Maroun. 2000. Evidence for an
interferon-related inflammatory reaction in the trisomy 16 mouse brain leading to caspase-1-mediated
neuronal apoptosis. J Neuroimmunol 110(1–2):66–75.
Hamilton, J. A. 2008. Colony-stimulating factors in inflammation and autoimmunity. Nat Rev Immunol 8
(7):533–44.
Hardy, J. A., and G. A. Higgins. 1992. Alzheimer’s disease: the amyloid cascade hypothesis. Science 256
(5054):184–85.
Heneka, M. T., M. J. Carson, J. E. Khoury, et al. 2015. Neuroinflammation in Alzheimer’s disease. The
Lancet Neurol 14(4):388–405.
Hoenen, C., Gustin, A., Birck, C., Kirchmeyer, M., Beaume, N., Felten, P., Grandbarbe, L., Heuschling,
P., Heurtaux, T. 2016. Alpha-Synuclein Proteins Promote Pro-Inflammatory Cascades in Microglia:
Stronger Effects of the A53T Mutant. PLoS One 11(9):e0162717.
Hohlfeld, R., and H. Wekerle. 2004. Autoimmune concepts of multiple sclerosis as a basis for selective
immunotherapy: from pipe dreams to (therapeutic) pipelines. Proc Natl Acad Sci U S A 101(Suppl 2):
14599–606.
Hristovska, I., and O. Pascual. 2016. Deciphering resting microglial morphology and process motility
from a synaptic prospect. Front Integrat Neurosci 9:73–73.
Hu, X., R. K. Leak, Y. Shi, et al. 2015. Microglial and macrophage polarization-new prospects for brain
repair. Nat Rev Neurol 11(1):56–64.
Itagaki, S., P. L. McGeer, H. Akiyama, S. Zhu, and D. Selkoe. 1989. Relationship of microglia and
astrocytes to amyloid deposits of Alzheimer disease. J Neuroimmunol 24(3):173–82.
Jakel, S., and L. Dimou. 2017. Glial cells and their function in the adult brain: a journey through the
history of their ablation. Front Cell Neurosci 11:24.
Kawanokuchi, J., T. Mizuno, H. Takeuchi, et al. 2006. Production of interferon-gamma by microglia.
Mult Scler 12(5):558–64.
Kawashima, T., K. Murata, S. Akira, et al. 2001. STAT5 induces macrophage differentiation of M1
leukemia cells through activation of IL-6 production mediated by NF-kappaB p65. J Immunol 167
(7):3652–60.
Kim, O. S., E. J. Park, E. H. Joe, and I. Jou. 2002. JAK–STAT signaling mediates gangliosides-induced
inflammatory responses in brain microglial cells. J Biol Chem 277(43):40594–601.
Kisseleva, T., S. Bhattacharya, J. Braunstein, and C. W. Schindler. 2002. Signaling through the JAK–
STAT pathway, recent advances and future challenges. Gene 285(1–2):1–24.
Kraft, A. D., L. S. Kaltenbach, D. C. Lo, and G. J. Harry. 2012. Activated microglia proliferate at neurites
of mutant huntingtin-expressing neurons. Neurobiol Aging 33(3):621.e17–33.
Krausgruber, T., K. Blazek, T. Smallie, et al. 2011. IRF5 promotes inflammatory macrophage polariza­
tion and TH1-TH17 responses. Nat Immunol 12(3):231–38.
Lang, R., D. Patel, J. J. Morris, R. L. Rutschman, and P. J. Murray. 2002. Shaping gene expression in
activated and resting primary macrophages by IL-10. J Immunol 169(5):2253–63.
Laurence, A., M. Pesu, O. Silvennoinen, and J. O’Shea. 2012. JAK kinases in health and disease: an
update. Open Rheumatol J 6:232–44.
Lee, D. Y., C. J. Hwang, J. Y. Choi, et al. 2017. KRICT-9 inhibits neuroinflammation, amyloidogenesis
and memory loss in Alzheimer’s disease models. Oncotarget 8(40):68654–67.
Lee, H. C., K. L. Tan, P. S. Cheah, and K. H. Ling. 2016. Potential role of JAK–STAT signaling pathway
in the neurogenic-to-gliogenic shift in Down syndrome brain. Neural Plast 2016:7434191.
Li, S., J. Sheng, J. K. Hu, et al. 2013. Ribonuclease 4 protects neuron degeneration by promoting
angiogenesis, neurogenesis, and neuronal survival under stress. Angiogenesis 16(2):387–404.
Li, X., L. Zhao, J.-J. Han, et al. 2018. Carnosol modulates Th17 cell differentiation and microglial switch
in experimental autoimmune encephalomyelitis. Front Immunol 9:1807.
98 JAK-STAT Signaling in Diseases

Liddelow, S. A., K. A. Guttenplan, L. E. Clarke, et al. 2017. Neurotoxic reactive astrocytes are induced by
activated microglia. Nature 541(7638):481–87.
Liebner, S., R. M. Dijkhuizen, Y. Reiss, K. H. Plate, D. Agalliu, and G. Constantin. 2018.
Functional morphology of the blood-brain barrier in health and disease. Acta Neuropathol
135(3):311–36.
Ling, K. H., C. A. Hewitt, K. L. Tan, et al. 2014. Functional transcriptome analysis of the postnatal brain
of the Ts1Cje mouse model for Down syndrome reveals global disruption of interferon-related
molecular networks. BMC Genomics 15:624.
Liu, C. Y., Y. Yang, W. N. Ju, X. Wang, and H. L. Zhang. 2018. Emerging roles of astrocytes in neuro­
vascular unit and the tripartite synapse with emphasis on reactive gliosis in the context of
Alzheimer’s disease. Front Cell Neurosci 12:193.
Liu, C., Y. Li, J. Yu, et al. 2013. Targeting the shift from M1 to M2 macrophages in experimental
autoimmune encephalomyelitis mice treated with fasudil. PLOS One 8(2):e54841.
Liu, X., J. Liu, S. Zhao, et al. 2016. Interleukin-4 is essential for microglia/macrophage M2 polarization
and long-term recovery after cerebral ischemia. Stroke 47(2):498–504.
Lively, S., and L. C. Schlichter. 2018. Microglia responses to pro-inflammatory stimuli (LPS,
IFNgamma+TNFalpha) and reprogramming by resolving cytokines (IL-4, IL-10). Front Cell
Neurosci 12:215.
Lobo-Silva, D., G. M. Carriche, A. G. Castro, S. Roque, and M. Saraiva. 2016. Balancing the immune
response in the brain: IL-10 and its regulation. J Neuroinflammation 13(1):297.
Lotfi, N., R. Thome, N. Rezaei, et al. 2019. Roles of GM-CSF in the pathogenesis of autoimmune
diseases: an update. Front Immunol 10:1265.
Luo, Y., and S. G. Zheng. 2016. Hall of fame among pro-inflammatory cytokines: interleukin-6 gene and
its transcriptional regulation mechanisms. Front Immunol 7:604.
Lyman, M., D. G. Lloyd, X. Ji, M. P. Vizcaychipi, and D. Ma. 2014. Neuroinflammation: the role and
consequences. Neurosci Res 79:1–12.
MacDonald, M. E., C. M. Ambrose, M. P. Duyao, et al. 1993. A novel gene containing a trinucleotide
repeat that is expanded and unstable on Huntington’s disease chromosomes. Cell 72(6):971–83.
Maries, E., B. Dass, T. J. Collier, J. H. Kordower, and K. Steece-Collier. 2003. The role of α-synuclein in
Parkinson’s disease: insights from animal models. Nat Rev Neurosci 4:727.
Mayo, L., S. A. Trauger, M. Blain, et al. 2014. Regulation of astrocyte activation by glycolipids drives
chronic CNS inflammation. Nat Med 20:1147.
Meares, G. P., Y. Liu, R. Rajbhandari, et al. 2014. PERK-dependent activation of JAK1 and STAT3
contributes to endoplasmic reticulum stress-induced inflammation. Mol Cell Biol 34(20):3911–25.
Miller, J. R., K. K. Lo, R. Andre, et al. 2016. RNA-Seq of Huntington’s disease patient myeloid cells
reveals innate transcriptional dysregulation associated with proinflammatory pathway activation.
Hum Mol Genet 25(14):2893–904.
Moore, K. W., R. de Waal Malefyt, R. L. Coffman, and A. O’Garra. 2001. Interleukin-10 and the
interleukin-10 receptor. Annu Rev Immunol 19:683–765.
Mount, M. P., A. Lira, D. Grimes, et al. 2007. Involvement of interferon-γ in microglial-mediated loss of
dopaminergic neurons. J Neurosci 27(12):3328–37.
Na, Y. J., J. K. Jin, J. I. Kim, E. K. Choi, R. I. Carp, and Y. S. Kim. 2007. JAK–STAT signaling pathway
mediates astrogliosis in brains of scrapie-infected mice. J Neurochem 103(2):637–49.
Nagele, R. G., M. R. D’Andrea, H. Lee, V. Venkataraman, and H. Y. Wang. 2003. Astrocytes accumulate
A beta 42 and give rise to astrocytic amyloid plaques in Alzheimer disease brains. Brain Res 971
(2):197–209.
Nakagawa, Y., and K. Chiba. 2015. Diversity and plasticity of microglial cells in psychiatric and
neurological disorders. Pharmacol Ther 154:21–35.
Natarajan, C., and J. J. Bright. 2002. Curcumin inhibits experimental allergic encephalomyelitis by
blocking IL-12 signaling through Janus kinase-STAT pathway in T lymphocytes. J Immunol 168
(12):6506–13.
Nateghi Rostami, M., M. Douraghi, A. Miramin Mohammadi, and B. Nikmanesh. 2012. Altered
serum pro-inflammatory cytokines in children with down’s syndrome. Eur Cytokine Netw 23
(2):64–67.
JAK-STAT Pathway and Neuroinflammation 99

O’Shea, J. J., and R. Plenge. 2012. JAK and STAT signaling molecules in immunoregulation and
immune-mediated disease. Immunity 36(4):542–50.
Olanow, C. W., and W. G. Tatton. 1999. Etiology and pathogenesis of Parkinson’s disease. Annu Rev
Neurosci 22:123–44.
Orihuela, R., C. A. McPherson, and G. J. Harry. 2016. Microglial M1/M2 polarization and metabolic
states. Br J Pharmacol 173(4):649–65.
Park, H. J., Oh, S. H., Kim, H. N., Jung, Y. J., Lee, P.H. 2016. Mesenchymal stem cells enhance α­
synuclein clearance via M2 microglia polarization in experimental and human parkinsonian
disorder. Acta Neuropathol 132(5):685–701.
Petrucelli, L., C. O’Farrell, P. J. Lockhart, et al. 2002. Parkin protects against the toxicity associated with
mutant α-synuclein: proteasome dysfunction selectively affects catecholaminergic neurons. Neuron
36(6):1007–19.
Probst, A., J. Ulrich, and P. U. Heitz. 1982. Senile dementia of Alzheimer type: astroglial reaction to
extracellular neurofibrillary tangles in the hippocampus. An immunocytochemical and
electron-microscopic study. Acta Neuropathol 57(1):75–9.
Przanowski, P., M. Dabrowski, A. Ellert-Miklaszewska, et al. 2014. The signal transducers Stat1 and
Stat3 and their novel target Jmjd3 drive the expression of inflammatory genes in microglia. J Mol
Med (Berl) 92(3):239–54.
Pugazhenthi, S., Y. Zhang, R. Bouchard, and G. Mahaffey. 2013. Induction of an inflammatory loop by
interleukin-1beta and tumor necrosis factor-alpha involves NF-kB and STAT-1 in differentiated
human neuroprogenitor cells. PLOS One 8(7):e69585.
Qin, H., J. A. Buckley, L. Xinru, et al. 2016. Inhibition of the JAK–STAT pathway protects against
α-synuclein-induced neuroinflammation and dopaminergic neurodegeneration. J Neurosci 36
(18):5144–59.
Qin, H., S. A. Niyongere, S. J. Lee, B. J. Baker, and E. N. Benveniste. 2008. Expression and
functional significance of SOCS-1 and SOCS-3 in astrocytes. J Immunol (Baltimore, MD: 1950)
181(5):3167–76.
Qin, H., W.-I. Yeh, P. De Sarno, et al. 2012. Signal transducer and activator of transcription-3/suppressor
of cytokine signaling-3 (STAT3/SOCS3) axis in myeloid cells regulates neuroinflammation. Proc
Natl Acad Sci 109(13):5004–5009.
Quintana, A., M. Muller, R. F. Frausto, et al. 2009. Site-specific production of IL-6 in the central nervous
system retargets and enhances the inflammatory response in experimental autoimmune
encephalomyelitis. J Immunol 183(3):2079–88.
Rawlings, J. S., K. M. Rosler, and D. A. Harrison. 2004. The JAK–STAT signaling pathway. J Cell Sci 117
(Pt 8):1281–83.
Robinson, A. P., C. T. Harp, A. Noronha, and S. D. Miller. 2014. The experimental autoimmune
encephalomyelitis (EAE) model of MS: utility for understanding disease pathophysiology and
treatment. Handbook Clin Neurol 122:173–89.
Ross, C. A., and S. J. Tabrizi. 2011. Huntington’s disease: from molecular pathogenesis to clinical
treatment. Lancet Neurol 10(1):83–98.
Rothaug, M., C. Becker-Pauly, and S. Rose-John. 2016. The role of interleukin-6 signaling in nervous
tissue. Biochim Biophys Acta 1863(6 Pt A):1218–27.
Rutschman, R., R. Lang, M. Hesse, J. N. Ihle, T. A. Wynn, and P. J. Murray. 2001. Cutting edge:
Stat6-dependent substrate depletion regulates nitric oxide production. J Immunol 166(4):2173–77.
Saijo, K., B. Winner, C. T. Carson, et al. 2009. A Nurr1/CoREST pathway in microglia and astrocytes
protects dopaminergic neurons from inflammation-induced death. Cell 137(1):47–59.
Sakai, K., T. Ohta, S. Minoshima, et al. 1995. Human ribosomal RNA gene cluster: identification of the
proximal end containing a novel tandem repeat sequence. Genomics 26(3):521–26.
Samoilova, E. B., J. L. Horton, B. Hilliard, T. S. Liu, and Y. Chen. 1998. IL-6-deficient mice are resistant
to experimental autoimmune encephalomyelitis: roles of IL-6 in the activation and differentiation of
autoreactive T cells. J Immunol 161(12):6480–86.
Sanchez-Ventura, J., J. Amo-Aparicio, X. Navarro, and C. Penas. 2019. BET protein inhibition regulates
cytokine production and promotes neuroprotection after spinal cord injury. J Neuroinflammation
16(1):124.
100 JAK-STAT Signaling in Diseases

Schindler, C., D. E. Levy, and T. Decker. 2007. JAK–STAT signaling: from interferons to cytokines. J Biol
Chem 282(28):20059–63.
Schulz-Schaeffer, W. J. 2010. The synaptic pathology of α-synuclein aggregation in dementia with
Lewy bodies, Parkinson’s disease and Parkinson’s disease dementia. Acta Neuropathologica
120(2):131–43.
Sherer, T. B. 2011. Biomarkers for Parkinson’s disease. Sci Trans Med 3(79):79ps14–79ps14.
Shukkur, E. A., A. Shimohata, T. Akagi, et al. 2006. Mitochondrial dysfunction and tau hyperpho­
sphorylation in Ts1Cje, a mouse model for Down syndrome. Hum Mol Genet 15(18):2752–62.
Silvestroni, A., R. L. Faull, A. D. Strand, and T. Moller. 2009. Distinct neuroinflammatory profile in
post-mortem human Huntington’s disease. Neuroreport 20(12):1098–103.
Simmons, D. A., M. Casale, B. Alcon, N. Pham, N. Narayan, and G. Lynch. 2007. Ferritin accumulation
in dystrophic microglia is an early event in the development of Huntington’s disease. Glia 55
(10):1074–84.
Simon, A. R., U. Rai, B. L. Fanburg, and B. H. Cochran. 1998. Activation of the JAK–STAT pathway by
reactive oxygen species. Am J Physiol 275(6 Pt 1):C1640–52.
Sospedra, M., and R. Martin. 2005. Immunology of multiple sclerosis. Annu Rev Immunol 23:683–747.
Streit, W. J., and M. B. Graeber. 1993. Heterogeneity of microglial and perivascular cell populations:
insights gained from the facial nucleus paradigm. Glia 7(1):68–74.
Streit, W. J., R. E. Mrak, and W. S. T. Griffin. 2004. Microglia and neuroinflammation: a pathological
perspective. J Neuroinflammation 1(1):14.
Su, F., F. Bai, and Z. Zhang. 2016. Inflammatory Cytokines and Alzheimer’s disease: a review from the
perspective of genetic polymorphisms. Neurosci Bull 32(5):469–80.
Subramaniam, S. R., and H. J. Federoff. 2017. Targeting microglial activation states as a therapeutic
avenue in Parkinson’s disease. Front Aging Neurosci 9:176.
Sullivan, K. D., H. C. Lewis, A. A. Hill, et al. 2016. Trisomy 21 consistently activates the interferon
response. Elife 5:e16220.
Tanabe, K., R. Matsushima-Nishiwaki, S. Yamaguchi, H. Iida, S. Dohi, and O. Kozawa. 2010. Mechanisms of
tumor necrosis factor-alpha-induced interleukin-6 synthesis in glioma cells. J Neuroinflammation 7:16.
Tanaka, S., A. Ishii, H. Ohtaki, S. Shioda, T. Yoshida, and S. Numazawa. 2013. Activation of microglia
induces symptoms of Parkinson’s disease in wild-type, but not in IL-1 knockout mice.
J Neuroinflammation 10:143.
Träger, U., A. Magnusson, N. L. Swales, et al. 2013. JAK–STAT signalling in Huntington’s disease
immune cells. PLOS Curr. 5:ecurrents.hd.5791c897b5c3bebeed93b1d1da0c0648.
Troncoso-Escudero, P., A. Parra, M. Nassif, and R. L. Vidal. 2018. Outside in: unraveling the role of
neuroinflammation in the progression of Parkinson’s disease. Front Neurol 9:860.
Tugal, D., X. Liao, and M. K. Jain. 2013. Transcriptional control of macrophage polarization. Arter­
ioscler Thromb Vasc Biol 33(6):1135–44.
von Bartheld, C. S., J. Bahney, and S. Herculano-Houzel. 2016. The search for true numbers of neurons and
glial cells in the human brain: a review of 150 years of cell counting. J Comp Neurol 524(18):3865–95.
Wan, J., A. K. Fu, F. C. Ip, et al. 2010. Tyk2/STAT3 signaling mediates beta-amyloid-induced neuronal
cell death: implications in Alzheimer’s disease. J Neurosci 30(20):6873–81.
Wang, W. Y., M. S. Tan, J. T. Yu, and L. Tan. 2015. Role of pro-inflammatory cytokines released from
microglia in Alzheimer’s disease. Ann Transl Med 3(10):136.
Weinstock-Guttman, B., R. M. Ransohoff, R. P. Kinkel, and R. A. Rudick. 1995. The interferons:
biological effects, mechanisms of action, and use in multiple sclerosis. Ann Neurol 37(1):7–15.
Wen, Y., W. H. Yu, B. Maloney, et al. 2008. Transcriptional regulation of beta-secretase by p25/cdk5 leads
to enhanced amyloidogenic processing. Neuron 57(5):680–90.
Wilcock, D. M., and W. S. Griffin. 2013. Down’s syndrome, neuroinflammation, and Alzheimer neuro­
pathogenesis. J Neuroinflammation 10:84.
Wilcock, D. M., J. Hurban, A. M. Helman, et al. 2015. Down syndrome individuals with Alzheimer’s
disease have a distinct neuroinflammatory phenotype compared to sporadic Alzheimer’s disease.
Neurobiol Aging 36(9):2468–74.
Wilhelmsson, U., E. A. Bushong, D. L. Price, et al. 2006. Redefining the concept of reactive astrocytes as
cells that remain within their unique domains upon reaction to injury. Proc Natl Acad Sci U S A 103
(46):17513–18.
JAK-STAT Pathway and Neuroinflammation 101

Wynn, T. A., A. Chawla, and J. W. Pollard. 2013. Macrophage biology in development, homeostasis and
disease. Nature 496(7446):445–55.
Xing-Long, G., C.-X. Long, L. Sun, C. Xie, X. Lin, and H. Cai. 2010. Astrocytic expression of Parkinson’s
disease-related A53T α-synuclein causes neurodegeneration in mice. Molecular Brain 3(1):12.
Xu, Y., M. Ikegami, Y. Wang, Y. Matsuzaki, and J. A. Whitsett. 2007. Gene expression and biological
processes influenced by deletion of Stat3 in pulmonary type II epithelial cells. BMC genomics 8:455.
Yan, Z., S. A. Gibson, J. A. Buckley, H. Qin, and E. N. Benveniste. 2018. Role of the JAK–STAT signaling
pathway in regulation of innate immunity in neuroinflammatory diseases. Clin Immunol 189:4–13.
Yoda, A., Y. Yoda, S. Chiaretti, et al. 2010. Functional screening identifies CRLF2 in precursor B-cell
acute lymphoblastic leukemia. Proc Natl Acad Sci U S A 107(1):252–57.
Zamanian, J. L., L. Xu, L. C. Foo, et al. 2012. Genomic analysis of reactive astrogliosis. J Neurosci
32(18):6391–410.
Zdaniuk, G., T. Wierzba-Bobrowicz, G. M. Szpak, and T. Stepien. 2011. Astroglia disturbances during
development of the central nervous system in fetuses with Down’s syndrome. Folia Neuropathol
49(2):109–14.
Zhang, J., F. Chen, W. Li, et al. 2012. 14-3-3ζ interacts with Stat3 and regulates its constitutive activation
in multiple myeloma cells. PLOS One 7(1):e29554.
Zhang, W., T. Wang, Z. Pei, et al. 2005. Aggregated alpha-synuclein activates microglia: a process leading
to disease progression in Parkinson’s disease. Faseb J 19(6):533–42.
Zhao, R., S. Xing, Z. Li, et al. 2005. Identification of an acquired JAK2 mutation in polycythemia vera.
J Biol Chem 280(24):22788–92.
7
JAK-STAT Signaling in Cardiovascular Disease

George W. Booz
Department of Pharmacology and Toxicology
The University of Mississippi Medical Center
Jackson, MS

Raffaele Altara
Institute for Experimental Medical Research
Oslo University Hospital and University of Oslo
Oslo, Norway
KG Jebsen Center for Cardiac Research
Oslo, Norway
Department of Pathology, School of Medicine
University of Mississippi Medical Center
Jackson, MS

Sean P. Didion
Department of Pharmacology and Toxicology
The University of Mississippi Medical Center
Jackson, MS
Department of Neurology
The University of Mississippi Medical Center
Jackson, MS

7.1 Introduction
Janus kinase (JAK) and signal transducer and activator of transcription (STAT) comprise a key
signaling pathway for a diverse array of stimuli, including a number of growth factors and
cytokines. The JAK family of tyrosine kinases includes JAK1, JAK2, JAK3, and tyrosine kinase
2 (Tyk2). The STAT family of transcription factors upon which the JAKs act include STAT1,
STAT2, STAT3, STAT4, STAT5A, STAT5B, and STAT6. Pharmacological inhibitors and
genetically modified JAK-STAT mice have yielded important insight into the functional role of
JAK-STAT signaling. While much of our knowledge of JAK-STAT signaling has been gained
from the role of JAK-STAT in the pathophysiology of immunodeficiency and cancer, the role of
JAK-STAT in the cardiovascular system in health and disease has only recently begun to
emerge. The goal of this chapter is to highlight the role of JAK-STAT in the cardiovascular
system in aging, atherosclerosis, diabetes, heart failure, hypertension, and stroke. We will focus
our discussion on the roles of JAK1 and JAK2, and STAT1 and STAT3, which account for the
majority of JAK and STAT activities in the cardiovascular system. Discussion will be primarily

103
104 JAK-STAT Signaling in Diseases

focused on the role of JAK-STAT signaling in the cardiovascular system, as well as some of the
specific ligands and receptors pathways that utilize JAK-STAT and the detrimental effects
associated with chronic activation of the JAK-STAT pathway. Areas of future research as well
as some potential therapeutic approaches to target JAK-STAT in cardiovascular disease will
also be briefly highlighted.

7.2 Canonical and Non-Canonical JAK-STAT Signaling


7.2.1 JAK and STAT Families
The JAK-STAT signaling pathway is an integral component of a number of cytokines and growth
factors. JAKs are non-receptor tyrosine kinases that contain a kinase and pseudokinase domain,
the latter of which plays an important regulatory role for the kinase domain. For that reason,
JAKs are named after Janus, the Roman God of gates and passages, who is symbolized with two
faces. Of the four mammalian JAK family members, JAK1 and JAK2, and to a lesser extent Tyk2
are expressed in cardiac and vascular cell types (Booz et al., 2002; Didion, 2017; Dutzmann et al.,
2015). JAKs are also characterized by seven highly conserved JAK homology (JH) domains. The
C-terminal region contains the JH1 domain with an autophosphorylation KE/DYY motif
required for activation, as well as the JH2 domain that regulates kinase activity and is a docking
site for STAT proteins. Originally thought to be a pseudokinase domain, JH2 has recently been
shown to be a dual-specificity protein kinase that negatively regulates JH1 activity, at least for
JAK2 (Hubbard, 2017; Ungureanu et al., 2011). The N-terminal regions are important for JAK-
receptor interactions. Once activated, the JAKs phosphorylate receptor tyrosine residues that
serve as docking sites to recruit STAT proteins, which are subsequently phosphorylated by the
JAK kinases. Tyrosine phosphorylated STATs dimerize via intermolecular Src homology (SH2)­
phosphotyrosine interactions, translocate to the nucleus, and initiate or suppress transcription of
STAT target genes. JAK-STAT-mediated transcription is tightly regulated by a number of
inhibitory processes including cytoplasmic and nuclear protein tyrosine phosphatases, as well as
expression of suppressor of cytokine signaling (SOCS) and protein inhibitor of activated STAT
(PIAS), both endogenous inhibitors of STAT (Chung et al., 1997; Krebs and Hilton, 2001; Shuai,
2006; Xu and Qu, 2008).
All seven mammalian STAT family members are reported to be expressed in both cardiac and
vascular cells: STAT1, STAT2, STAT3, STAT4, STAT5a, STAT5b, and STAT6 (Booz et al.,
2002). All STAT family members, except for STAT2 and STAT6 (and STAT3β), contain a serine
residue within a conserved PMSP or PSP motif of the transcriptional activation domain (TAD),
which serves a regulatory function. For instance, Ser727 in both STAT1 and STAT3 enhances its
transcriptional activity, but may also participate in terminating DNA binding by recruiting
a tyrosine phosphatase. It has also been reported that serine phosphorylated STAT1 and STAT3
may regulate gene expression independently of tyrosine phosphorylation (Booz et al., 2002).
Certain G protein-coupled receptors, such as the angiotensin II type 1 receptor (AT1R), also
signal, in part, through JAK-STAT (Booz et al., 2002). Tyrosine kinase and G protein-coupled
receptors may also use Src kinase to phosphorylate STAT proteins depending upon cell type and
the particular STAT isoform involved.

7.2.2 JAK-STAT Stimuli and Inhibitors


JAK-STAT signaling is activated by a large number of cytokines and growth factors that have
important effects on the cardiovascular system, including interferon-α (IFN-α), IFN-γ, IL-3, IL-5,
and IL-6 as well as angiotensin II and thrombin, as some respective examples (Darnell et al.,
1994; Didion, 2017; Guschin et al., 1995; Kaptein et al., 1996; Madamanchi et al., 2001; Marrero,
2005; Marrero et al., 1995; Okuda et al., 1999) (Figure 7.1). In addition, mechanical stretch has
JAK-STAT Signaling in Cardiovascular Disease 105

been shown to activate JAK-STAT (Pan et al., 1999). Early studies using the specific JAK2
inhibitor, AG490, provided some of the first evidence for JAK-STAT signaling in the response of
a number of cardiovascular cells types to various cytokines. However, AG490 is not particularly
specific and is also a potent inhibitor of EGF receptor and ErbB2 tyrosine kinases (Booz et al.,
2002). In light of the broad nature of JAK signaling, and difficulties surrounding the development
of JAK isoform-specific inhibitors, development of JAK-STAT inhibitors has primarily focused
on targeting STAT3. A number of small molecular inhibitors targeting the SH2 domain of STAT3
have been developed, including STATTIC and S31-201 (Schust et al., 2006; Siddiquee et al.,
2007). These compounds disrupt STAT3 dimerization and decrease its DNA-binding and tran­
scriptional activity albeit through slightly different mechanisms. One of the first-generation
inhibitors of this kind was STATTIC (Schust et al., 2006). S3I-201, which is fairly specific for
STAT3, is perhaps the most studied inhibitor of this kind recently. Various analogs of S31-201 of
increased potency have been developed and some have entered clinical trials. However, since
STAT3 may translocate to the nucleus and bind DNA without Y705 phosphorylation, (Zouein
et al., 2015) targeting the SH2 domain has met with limited success due to the inability of current
inhibitors to completely inhibit STAT3 activity.
For the IL-6 family of cytokines, signaling is terminated by recruitment of SOCS3 to the
receptor subunit close to where the JAK is located, after phosphorylation of the docking
residues by the activated JAKs (Kurdi and Booz, 2007) (Figure 7.2). SOCS3 inhibits JAK
activity through its kinase inhibitory domain and, unlike SOCS1, shows low affinity for JAK1
or JAK2 in the absence of receptor binding. SOCS3 and SOCS1 are induced by IL-6-type
cytokines downstream of STAT3 and/or STAT1, and thus serve as endogenous negative
feedback inhibitors. Besides directly inhibiting JAK kinase activity, SOCS1 and SOCS3 may
also terminate IL-6-type cytokine signaling by facilitating JAK2 ubiquitination and subsequent
proteasomal degradation (Kile and Alexander, 2001). Ablation of cardiac SOCS3 appears to be
protective against ischemia-reperfusion injury in the mouse by enhancing STAT3 and other
protective signaling mechanisms (Nagata et al., 2015; Oba et al., 2012). On the other hand,
SOCS3 deletion in cardiomyocytes was reported to aggravate DOCA-salt-induced hypertrophic

FIGURE 7.1 JAK-STAT signaling in the cardiovascular system is activated by a number of cytokines, such as IL-6,
and growth factors, such as angiotensin II. Mechanical stretch, which is a functional component of cardiac and
vascular muscle, also involves JAK-STAT signaling and target gene transcription. JAK-STAT signaling has been
shown to contribute to cardiovascular disease, in part, by alterations in gene expression that contribute to a number of
processes and phenotypes, including that shown above, but most particularly, endothelial dysfunction, immune cell
activation, and cardiovascular hypertrophy.
106 JAK-STAT Signaling in Diseases

FIGURE 7.2 Interleukin-6 signaling via JAK-STAT in cardiovascular cells involves JAK1 and STAT3. IL-6R lacks
intrinsic signaling capability; however, association of the IL-6R with the transmembrane protein, gp130, which
contains docking site for the tyrosine kinase JAK1, mediates IL-6 signaling. JAK1 is constitutively associated with
gp130 and upon IL-6 ligand/receptor interaction induces autophosphorylation and activation of JAK1. JAK1 then
phosphorylates specific tyrosines residues within the cytoplasmic portion of gp130, which function as docking sites for
STAT3 and phosphorylation of STAT3 by JAK1. Phosphorylated STAT3 dimerize, then translocates to the nucleus to
increase or decrease transcription of STAT3 target genes. STAT3-induced transcription can be inhibited by SOCS3,
PIAS, as well as cytosolic or nuclear phosphatases. STAT3-mediated transcription can also be inhibited pharmacolo­
gically by the small molecule inhibitors, STATTIC and S3I-201. IL-6, Interleukin-6; IL-6R, Interleukin-6 Receptor;
JAK, Janus-associated Kinase; STAT, Signal Transducer and Activator of Transcription; SOCS, Suppressor Of
Cytokine Signaling; PIAS, Protein Inhibitor of Signaling; PTPs, protein tyrosine phosphatases.

cardiac remodeling and was associated with age-related cardiomyopathy and ventricular
arrhythmias; although a non-specific effect of prolonged cre-recombinase expression cannot
be discounted (Liu et al., 2018; Pugach et al., 2015; Yajima et al., 2011). The harmful effects of
SOCS3 deletion have been attributed to unbridled inflammatory signaling downstream of the
IL-6 type cytokines.

7.2.3 JAK-STAT Transgenic/Knockout Mice


While pharmacological inhibitors of JAK-STAT have provided great insight into the role of JAK­
STAT in the transduction of growth factor and cytokine signaling the development and avail­
ability of JAK- and STAT-deficient mice has revealed non-redundant roles of specific JAKs and
STATs in embryonic development as well as signaling transduction and gene transcription. Global
JAK- and STAT-deficient mouse models have been developed for each of the JAK and STAT
isoforms (Baird et al., 1998; Durbin et al., 1996; Feldman et al., 1997; Kaplan et al., 1996a,
1996b; Liu et al., 1997; Meraz et al., 1996; Neubauer et al., 1998; Nosaka et al., 1995; Parganas
et al., 1998; Park et al., 2000; Rodig et al., 1998; Shimoda et al., 1996; Smit et al., 1997; Takeda
et al., 1997; Thomis et al., 1995; Udy et al., 1997). The discussion here, however, will be limited to
those JAK and STAT isoforms most relevant to the cardiovascular system, namely JAK1, JAK2,
JAK3, STAT1, and STAT3.
JAK-STAT Signaling in Cardiovascular Disease 107

Global deficiency of JAK1 is perinatal lethal (Rodig et al., 1998). Despite the lethality associated
with global JAK1 deficiency, JAK1-/- embryos appear grossly normal despite weighing 40% less than
wild-type mice. JAK1-/- mice display normal erythrocyte, mononuclear, neutrophil, and platelets
numbers (Rodig et al., 1998). JAK1-/- mice are, however, characterized by severely reduced B cell
and thymocyte cell numbers reflective of primary defects in B cell differentiation and thymocyte
production, respectively. Responses to type II cytokines, such as IFN-α and IFN-γ, are completely
absent in JAK1-/- deficient cells, suggesting a selective role of JAK1 in IFN-α and IFN-γ signaling
(Rodig et al., 1998). In contrast, responses to IL-6 in JAK1-/- deficient immune cells are diminished,
but not completely abolished (Rodig et al., 1998). In order to circumvent some of the issues
associated with lethality of global JAK1 deficiency, Cre-recombinase conditional JAK1-deficient
mice have been developed (Sakamoto et al., 2016).
Global JAK2 deficiency is associated with lethality around embryonic days 10–13. Global
JAK2-/- mice are characterized by severe anemia and the absence of red blood cells in the fetal
liver, thus, it has been concluded that the lethality associated with loss of JAK2 expression is due
to the lack of definitive erythropoiesis (Neubauer et al., 1998; Parganas et al., 1998). Global
JAK2-/- mice display deficits in cytokine responses to the type II cytokine IFN-γ and no response
to cytokines that utilize beta chain receptor signaling, such as thrombopoietin, IL-3, IL-5, and
GM-CSF (Neubauer et al., 1998; Parganas et al., 1998). In contrast, cellular responses to IFN-α,
IFN-γ, and IL-6 appear normal. Other than the absence of red blood cells, JAK2-/- mice display
normal immune cell numbers, suggesting that JAK2 is not required for the development of
lymphoid lineage progenitors. In order to circumvent the issue of lethality associated with global
JAK2-deficiency, a number of conditional and cell-specific (including a cardiac-, endothelial-, and
smooth muscle-cell specific) JAK2-deficient mouse models have been developed (Frenzel et al.,
2006; Gan et al., 2015; Kirabo et al., 2011a, 2011b; Krempler et al., 2004; Park et al., 2013; Yang
et al., 2013). For example, cell–specific-JAK2 knockout models have been used to investigate the
role of JAK2 signaling in response to angiotensin II-induced hypertension and endothelial
dysfunction (Kirabo et al., 2011a, 2011b).
Unlike JAK1 or JAK2 deficiency, JAK3-deficient mice are viable; however, they are character­
ized by a severe combined immune deficiency (SCID) phenotype (Nosaka et al., 1995; Thomis
et al., 1995). The SCID-like phenotype is due to the fact that the expression of JAK3 is primarily
restricted to immune cells, whereas JAK1 and JAK2 are ubiquitously expressed. The SCID-like
phenotype of JAK3-deficient mice is primarily due to defects in IL-2, IL-4, and IL-7 signaling
both of which utilize the common gamma chain receptor. Examination of T cell development
suggests that discrete defects at the CD4-CD8-stage of T cell maturation are responsible for the
loss of immature T cells through apoptosis (Baird et al., 1998). The lack of B cell precursors in
bone marrow and the spleen and the lack of mature T cells highlight the essential role of JAK3 in
the development of lymphoid cells.
Global STAT1 deficient mice are viable and demonstrate no overt developmental abnormalities;
however, Stat1-/- mice are highly susceptible to pathogenic infection (Durbin et al., 1996; Meraz
et al., 1996). Stat1-/- mice are completely nonresponsive to IFN-α and IFN-γ indicating that STAT1
is indispensable for interferon signaling. In contrast, Stat1-/- mice are characterized by normal, or
near normal, responses to leukemia inhibitory factor, IL-6, platelet-derived growth factor, and
epidermal growth factor suggesting that there is functional redundancy for these cytokines and
growth factors, most likely provided for by STAT3. Interestingly, STAT1 and STAT3 show similar
DNA-binding specificity and distinct patterns of binding to various transcriptional enhancer
elements. Cardiac cell-specific STAT1 deficiency limits cardiac damage in response to myocardial
infarction, in part, by enhancing autophagy (McCormick et al., 2012).
In contrast to STAT1, global STAT3 deficiency is lethal with fetal demise occurring at
embryonic days 6–8 (Takeda et al., 1997). Thus, STAT3 expression is essential for early embryonic
development. Cell-specific STAT3 deficient mice have been developed to circumvent the issue of
embryonic lethality associated with the complete absence of STAT3 (Jacoby et al., 2003; Kano
et al., 2003). For example, cardiac specific Stat3-/- mice have been found to be more sensitive
to doxorubicin-induced heart failure, lipopolysaccharide-induced cardiac cell death, and an
108 JAK-STAT Signaling in Diseases

increased incidence of heart failure with normal aging, suggesting that STAT3 may play
a protective role in the heart (Jacoby et al., 2003). Transgenic expression of a mutant STAT3
gene, SA/SA mice in which Ser727 phosphorylation site has been mutated, is associated with an
approximately 50% reduction in STAT3 activity, as phosphorylation of Ser727 is required for
maximal STAT3 transcriptional activity (Shen et al., 2004). SA/SA mice have been utilized to
understand the role of angiotensin II in cardiac hypertrophy as discussed later on below (Zouein
et al., 2013a).

7.3 JAK-STAT in Cardiovascular Aging


Aging is associated with an increased incidence of cardiovascular disease, including atherosclero­
sis, vascular cognitive impairment, hypertension, and stroke (Buford, 2016; Lo Coco et al., 2016;
Wen and Wong, 2018). Blood pressure also increases with age at a rate of approximately 7 mmHg
per decade of life over the age of 40 (Wolf-Maier et al., 2003). Interestingly, blood pressure has
been found to correlate positively with increases in plasma IL-6 levels in healthy individuals (Chae
et al., 2001). Moreover, plasma levels of IL-6 have been found to be significantly higher in aged as
compared to young individuals (Maggio et al., 2006). Increases in plasma IL-6 are associated with
increases in cardiovascular- and all-cause mortality in elderly populations (Li et al., 2017). In
contrast, IL-2 levels have been found to decrease with age (Fulop et al., 2006). Thus, in addition
to a global reduction in immune cell responses to pathogens, it appears that aging is also
characterized by chronic, low-grade inflammation, and alterations in cytokine levels, particularly
alterations in levels of cytokines that signal via JAK-STAT, such as IL-6.
Alterations in immune cell number and activity are altered with aging, such that cytokine-specific
responses are reduced in geriatric patients (Longo et al., 2012; Shen-Orr et al., 2016). For example,
aging is associated with reductions in T cell responses to IL-6, IFN-α, IFN-γ, and IL-21 and
a diminished ability of these cytokines to phosphorylate STAT1, STAT3, and STAT5. (Shen-Orr
et al., 2016). While the ability of cytokines to phosphorylate STAT is diminished with aging, this
appear to be reflective of the fact that basal levels of phosphorylated STAT are increased in aging,
due to a more chronic stimulation of JAK-STAT signaling. Initial reports suggest that inflammatory
cytokines, such as IL-6, and STAT1 and STAT3 expression and phosphorylation are increased with
age (O’Brown et al., 2015). It is this higher level of phosphorylated STAT that has been suggested
to contribute to the greater incidence of atherosclerosis and diastolic dysfunction in aged humans
(Shen-Orr et al., 2016). Interestingly, the response to pathogens is greatly diminished with aging,
due, in part, to the fact that basal levels of phosphorylated STAT are higher with aging. Thus, the
loss of the robustness of the response to cytokines that signal through JAK-STAT may be a major
reason why older individuals are more prone to infection.
Despite the fact that JAK-STAT plays a major role in the regulation of immune cells, very little
is known regarding the specific alterations that occur in the various immune cell populations with
aging and cardiovascular disease. Interestingly, inhibition of JAK signaling has been shown to
alleviate cellular senescent phenotypes and frailty commonly associated with aging (Xu et al.,
2015). Although pharmacological inhibitors of JAK have provided some insight into the role of
JAK-STAT signaling in aging, the role of JAK-STAT signaling in the cardiovascular system with
age has not been fully explored and is thus an area that requires further study.

7.4 JAK-STAT in Atherosclerosis


JAK-STAT signaling has been implicated in various aspects of atherosclerosis in a temporal and
cellular context. Principally, chronic activation of JAK-STAT signaling contributes to vascular
inflammation and atherosclerosis initiation and progression. In endothelial cells from different
vascular beds for instance, the IL-6 type cytokines induce via STAT3 the expression of various
JAK-STAT Signaling in Cardiovascular Disease 109

surface adhesion molecules, chemokines, and cytokines that are important for monocyte/macrophage
recruitment (Alturkmani et al., 2012; van Keulen et al., 2018). Induction of SOCS3 in vascular
endothelial cells by flavonoids by an ERK1/2-mediated pathway was shown to block IL-6-induced
pro-inflammatory signaling (Wiejak et al., 2013). Endothelial cell STAT3 has also been implicated in
atherosclerotic lesion development due to the critical importance of STAT3 in vascular cell prolifera­
tion and angiogenesis (Li et al., 2018). Lastly, the inflammatory response of platelets to oxidized low-
density lipoprotein (ox-LDL) may involve activation of nuclear factor kappa B (NF-κB) and STAT3
(Sun et al., 2018).
STAT3 activation has been implicated as well in the proliferation, hypertrophy, and migration
of vascular smooth muscle (Chen et al., 2014; Kirchmer et al., 2014; Wang et al., 2016). The
dispersion of intimal smooth muscle cells may contribute to pathologic intimal thickening in
human atherosclerosis (Nakagawa and Nakashima, 2018). STAT3 has also been observed to
contribute to osteoblastic differentiation of vascular smooth muscle as seen in atherosclerosis,
while STAT1 may play a role in vascular smooth muscle de-differentiation (Kakutani et al., 2015;
Kirchmer et al., 2014).
Macrophages are key components of atheroma development. The role of STAT3 in macrophage
phenotypes is complicated and not fully understood. Evidence that the IL-6 family member
cardiotrophin-like cytokine factor 1 (CLCF1) increases macrophage-foam cell transition was
recently implicated in this process with ruxolitinib, a JAK inhibitor (Pasquin et al., 2018).
Accumulation of fat-laden pro-inflammatory M2 macrophages occurs with plaque development
and may lead to a necrotic core that, with plaque rupture, forms a thrombus and precipitates an
embolism. STAT3 activation was further implicated in the recruitment of additional macrophages
through their formation of chemokine (C-C motif) ligand 2 (CCL2), aka monocyte chemoattrac­
tant protein 1 (MCP1) (Zhao et al., 2018). Minimally modified LDL-induction of CCL2 was also
shown to involve enhancement of STAT3 promoter activity by mTORC1 (Ai et al., 2014). On the
other hand, evidence was recently reported to support the conclusion that the anti-atherogenic
actions of high-density lipoprotein (HDL) are attributable in part to sphingosine 1-phosphate
(S1P)-mediated inhibition of macrophage apoptosis through induction of survivin expression by
STAT3 (Feuerborn et al., 2017). Taken together, evidence suggests that JAK-STAT signaling
within vascular cells and immune cells are key components of the atherosclerotic processes related
to lesion development in the vascular wall.

7.5 JAK-STAT in Diabetes


With type 1 diabetes, levels and activity of STAT3 are reported to be reduced in cardiac and
vascular cell types (Das et al., 2015; Drenger et al., 2011; Owais et al., 2015). As a consequence,
the effectiveness of protective mechanisms such as ischemic post-conditioning in the heart is less
effective (Drenger et al., 2011). In addition, both hyperglycemia and the generation of reactive
oxygen species may impact ischemic post-conditioning and STAT3 activation by impairing
adiponectin/caveolin-3 signaling in the heart (Li et al., 2016a; Wang et al., 2013). Others have
implicated a loss of both PI3K/Akt and JAK2/STAT3 signaling due to increased cardiac PTEN
activity, as the basis for ablation of post-conditioning in diabetic hearts (Xue et al., 2016). In
diabetic leptin receptor null hearts (a model of type II diabetes), inhibition of mTORC1 with
rapamycin, an inhibitor of mTOR, was reported to offer some protection from ischemia­
reperfusion injury by restoring activation of STAT3, as well as Akt phosphorylation (Das et al.,
2015). The critical importance of STAT3 in the actions of rapamycin was further demonstrated
with a cardiomyocyte-specific STAT3 knockout model fed a high fat diet (Das et al., 2015). While
reduced STAT3 activation is implicated in compromised protection of the diabetic heart from
ischemia-reperfusion injury, a number of preclinical studies have implicated increased STAT3
activation in cardiac fibrosis in both type 1 and 2 diabetes (Chang et al., 2016; Fiaschi et al., 2014;
Lo et al., 2017; Wang et al., 2015).
110 JAK-STAT Signaling in Diseases

7.6 JAK-STAT in Heart Failure


In broad terms, heart failure may be classified into two basic subtypes depending upon whether
a primary effect on systolic function is involved. Heart failure with reduced ejection fraction
(HFrEF) is traditionally associated with this condition, whereas heart failure with preserved
ejection fraction (HFpEF) is a less understood disorder that is characterized by diastolic
dysfunction or increased stiffness of the myocardium. Two major causes of HFrEF are myocardial
infarction (MI) and cardiac hypertrophy secondary to hypertension. With HFrEF, the left
ventricle adopts a more elongated architecture with thinning wall thickness, which proves
ineffective in pumping blood. The STATs, in particular STAT3, have been implicated in both
cardiac protection from ischemia-reperfusion injury and somewhat paradoxically hypertension-
induced cardiac injury. Thus, STAT3 may play dual roles in cardiac protection and injury reflecting,
in part, the temporal nature of STAT3 activation in response to ischemia.
Evidence supports a central role for STAT3 in protecting the heart from ischemia-reperfusion
injury (Kurdi and Booz, 2007; Zouein et al., 2013b). Protection occurs either directly or as
a component of various forms of pre- or post-conditioning (i.e., ischemic, pharmacological, direct,
and remote). For instance, cardiac myocyte-specific STAT3-deficiency was found to have no affect on
myocardial infarct size, although ischemic and various forms of pharmacological preconditioning
were eliminated with the loss of STAT3 (Smith et al., 2004). On the other hand, larger infarct sizes and
greater apoptosis were observed 24 h after reperfusion in hearts of cardiac myocyte-specific Stat3-/­
mice, as well as impaired fractional shortening and dilated cardiomyopathy 7 days post-ischemia
(Hilfiker-Kleiner et al., 2004). STAT3 deficiency was also associated with enhanced mortality and
increased mRNA expression of pro-apoptotic/autophagy protein BNIP3 after 24 h of reperfusion,
whereas expression levels of the pro-survival gene HSP70 were reduced (Hilfiker-Kleiner et al., 2004).
Evidence was found that Stat3-/- myocyte VEGF expression, which is pro-angiogenic and linked to
STAT3 overexpression or activation in the heart, was not affected by loss of STAT3.
STAT3 has been implicated in the early short-lived phase of preconditioning, which does not
involve gene expression (Zouein et al., 2013b). In an inducible and cardiac-specific model of
STAT3 deficiency, loss of STAT3 expression was associated with upregulation of cardioprotective
(HO-1 and COX-2) and anti-apoptotic (e.g., Bcl-xL, Mcl-1, c-FLIPL, and c-FLIPS) genes in
response to delayed cardiac preconditioning (Bolli et al., 2011). With a similar model, cardio­
myocyte STAT3 deficiency was found to contribute to remodeling in the subacute phase of MI
(Enomoto et al., 2015). STAT3 loss from 11 to 24 days post-MI was associated with impaired
cardiac function and increased mortality. Increased expression of fibrosis-related genes and
worsened cardiac fibrosis were seen, likely due to enhanced death of cardiomyocytes. There was
also increased cardiac hypertrophy after MI with a reduction in capillary density.
A critical protective role for STAT3 in left ventricular remodeling from MI is also supported by
a study examining the effect of cardiac-specific deletion of SOCS3, the endogenous feedback inhibitor
of STAT3 signaling (Oba et al., 2012). Infarct size, myocardial apoptosis, and cardiac fibrosis were
remarkably reduced in these mice 14 days after MI. STAT3 activation and the expression of
antioxidants HO-1 and SOD2 were enhanced by SOCS3 deletion. Cardiac function and mortality
were improved. Similarly, reduced myocardial injury was observed 24 h after ischemia–reperfusion in
SOCS3 knockout mice (Nagata et al., 2015). This was associated with enhanced expression of the
STAT3 gene target and anti-apoptotic Bcl-2 family member, myeloid cell leukemia-1 (Mcl-1).
However, excessive activation of STAT3 in cardiac myocytes due to a mutant gp130 receptor was
associated with a worse outcome in subacute MI (Hilfiker-Kleiner et al., 2010). This was attributed to
increased inflammation; however, the contribution of other signaling events associated with gp130
receptor activation need to be taken into consideration (Zgheib et al., 2012).
Ischemia or ischemia-reperfusion is reported to induce the translocation of STAT3 to mitochondria
in the heart by as yet undefined pathways (Szczepanek et al., 2011). Overexpression of mitochondrial­
targeted STAT3 with a mutation in the DNA-binding domain was found to reduce reactive oxygen
species production and preserve mitochondrial function during ischemia (Szczepanek et al., 2011).
JAK-STAT Signaling in Cardiovascular Disease 111

STAT3 overexpression in mitochondria was also associated with higher survival and improved
contractile function, due in part to attenuation of mitochondrial permeability transition pore open­
ing, and decreased MI size with ischemia-reperfusion (Szczepanek et al., 2015).
Cardiac hypertrophy is a classical physiological response to chronic increases in arterial pressure.
Although initially beneficial in reducing wall stress, cardiac hypertrophy can progress to HFrEF (Booz,
2005; Drazner, 2011). Indeed, hypertension-induced left ventricular hypertrophy is an independent
predictor of morbidity and mortality. Early studies indicated that activation of STAT3 played an
important role in hypertrophy of cardiac myocytes. The IL-6 family of cytokines have been found to
induce the growth of isolated neonatal rat ventricular myocytes via STAT3 (Wollert et al., 1996). In
contrast, deficiency of IL-6 was found to have no effect on cardiac hypertrophy and function in the
transverse aortic constriction of pressure overload (Lai et al., 2012; Zhao et al., 2016). More recently,
deficiency of IL-6 was found to inhibit cardiac inflammation, fibrosis, and dysfunction in hypertension
produced by high-salt plus infusion of supraphysiological concentrations of angiotensin II (Yue et al.,
2010). Thus, the contribution of JAK-STAT to cardiac hypertrophy may be different depending on the
temporal activation of JAK-STAT signaling and on the specific etiology of hypertension.
Transgenic overexpression of STAT3 in mouse hearts was associated with cardiac hypertrophy
(Kunisada et al., 2000). In addition, the extent of AT1R-induced cardiac hypertrophy was
correlated with the accumulation of unphosphorylated STAT3 in the nucleus (Yue et al., 2010).
Unphosphorylated STAT3 is postulated to result from STAT3-induced STAT3 expression down­
stream of cytokine-induced STAT3 activation. Unphosphorylated STAT3, in turn, was associated
with the increased expression of a subset of cardiac hypertrophy related genes, such as osteopontin
and regulator of G protein signaling 2 (RGS2). Acetylation of Lys-685 may be important for
stabilizing the homodimers of unphosphorylated STAT3 in the absence of Tyr-705 phosphorylation
(Dasgupta et al., 2014). Thus, both the levels and localization of unphosphorylated versus
phosphorylated STAT3 may dictate the response to hypertrophic stimuli.
On the other hand, non-cardiomyocyte STAT3 may play a role in cardiac hypertrophy. An
earlier study on angiotensin II infusion of mice with reduced STAT3 activity due to a global
S277A mutation found that cardiac hypertrophy was attenuated along with an increase in
reparative patches of fibrosis (Zouein et al., 2013a). Moreover, cardiac hypertrophy may actually
be attenuated by endogenous STAT3. IL-10 was found to block isoproterenol- and transverse
aortic constriction (TAC)-induced cardiac hypertrophy, due to STAT3-mediated inhibition of NF­
κB and p38 activation (Verma et al., 2012). Left ventricular dysfunction, fibrosis, hypertrophy,
and fetal gene expression were found to be greater in IL-10 knockout mice.
In addition to an increase in the size of cardiac myocytes, pathological cardiac remodeling entails
increased fibrosis. STAT3 has been implicated in collagen synthesis and cardiac fibrosis. For instance,
hyper-activated STAT3 was observed in the myofibroblasts of mouse hearts with chronic β-adrenergic
stimulation (Zhang et al., 2016). EphrinB2, a cell surface transmembrane ligand, was recently found to
have pro-fibrotic actions in cardiac fibroblasts by a synergistic activation of STAT3 and Smad3 (Su
et al., 2017). With HFpEF, increased stiffness of both myofilaments and extracellular matrix is thought
to impair diastolic function (Paulus and Tschope, 2013; Tschope and Van Linthout, 2014; Zouein et al.,
2013a). The former is postulated to result from reduced protein kinase G (PKG)-mediated phosphor­
ylation of titin, (Paulus and Tschope, 2013; Tschope and Van Linthout, 2014; Zile et al., 2015) the
protein that determines passive elasticity of cardiomyocytes, and the latter from increased collagen
deposition and cross-linking (fibrosis) (Tschope and Van Linthout, 2014; Zile et al., 2015; Zouein et al.,
2013a). However, the role of STAT3 in the progression to HFpEF is largely unexplored.

7.7 JAK-STAT in Hypertension


Hypertension is a major risk factor for cardiovascular and cerebrovascular disease and events,
such as carotid artery disease, vascular dementia, myocardial infarction, and stroke (Boehme
et al., 2017; Jusufovic et al., 2016; Nagai et al., 2010; Rosendorff et al., 2015; Sztriha et al., 2009).
112 JAK-STAT Signaling in Diseases

Hypertension is also a major cause of cardiovascular-related morbidity and mortality (Patel et al.,
2016). Although pharmacological and genetic approaches to limit JAK-STAT signaling in
hypertension have been found to lower blood pressure and improve endothelial function in
preclinical models, the role of JAK-STAT in human hypertension has yet to be fully defined.
Angiotensin II, the major effector of the renin-angiotensin system (RAS), is a potent vasocon­
strictor and plays a major role in the development and maintenance of the hypertension
phenotype (Hussain and Awan, 2018). Plasma Angiotensin II levels are elevated in several forms
of hypertension, such as renovascular hypertension. This is most evident clinically in the utility of
pharmacological inhibitors of RAS signaling namely, angiotensin converting enzyme inhibitors,
AT1R blockers (such as losartan), and direct renin inhibitors in lowering blood pressure and
limiting the negative cardiovascular effects associated with hypertension irrespective of the
etiology (Te Riet et al., 2015). The majority of angiotensin II’s effects on the cardiovascular
system are mediated, in large part, through AT1R (Kawai et al., 2017; Oliverio et al., 1997; Ryan
et al., 2004). AT1R is a G-protein coupled receptor and ATR1 activation has been shown to
promote hypertrophy of cardiac and vascular muscle and fibrosis. The mitogenic and fibrotic
effects of angiotensin have been shown to contribute directly to impairment of cardiovascular
function (Booz et al., 2002; Chen et al., 2017; Didion and Faraci, 2003; Didion et al., 2000, 2005,
2009; Gomolak and Didion, 2014; Johnson et al., 2013; Schrader et al., 2007; Zouein et al.,
2013a). AT1R signaling occurs through multiple pathways, such as c-Src, ERK1/2, MAPK, PLC,
as well as JAK-STAT (Harvey et al., 2016; Marrero et al., 1997; Mehta and Griendling, 2007).
Activation of JAK-STAT by AT1R occurs in a unique manner, which will be discussed below, to
that of classical tyrosine kinase receptor signaling normally associated with canonical activation
of JAK-STAT signaling by cytokine receptors.
In seminal studies performed in the mid-1990s, angiotensin II was shown to rapidly increase
(within minutes) tyrosine phosphorylation of JAK2, but not JAK1, in cultured vascular smooth
muscle cells (Marrero et al., 1995, 1996). In subsequent studies, COS-7 cells (which normally
express high levels of STAT1) transfected with wild-type AT1R and JAK2 was found to produce
rapid and dose-dependent activation of JAK2 (i.e., JAK2 autophosphorylation) followed by
JAK2-dependent phosphorylation of STAT1 in response to angiotensin II (Ali et al., 2000). In
contrast, STAT1 phosphorylation was absent in angiotensin II-treated COS-7 cells transfected
with a catalytically inactive (dominant negative) JAK2, suggesting that JAK2 kinase activity is
required for STAT1 phosphorylation (Ali et al., 2000). Moreover, the effect of angiotensin II on
JAK2 activation and STAT1 phosphorylation were abrogated by the AT1R receptor blocker,
losartan, providing pharmacological evidence that angiotensin II-induced AT1R activation was
necessary for phosphorylation and activation of both JAK2 and STAT1 (Ali et al., 2000).
Subsequent to these studies, angiotensin II was also shown to stimulate phosphorylation of
STAT3 indirectly via angiotensin II-induced increases in IL-6 (Bhat and Baker, 1997; Didion,
2017; McWhinney et al., 1997; Recinos et al., 2007).
A major question to emerge from these early studies was: How exactly does AT1R promote
phosphorylation of JAK2? Evidence already indicated that AT1R did not possess intrinsic tyrosine
kinase activity. While studies clearly demonstrate that angiotensin II binding to the AT1R leads to
JAK2 activation, it was not entirely clear how this occurs; however, AT1R-dependent phospholipase
C activation leading to increases in intracellular calcium appears to be important for JAK2 autopho­
sphorylation (Frank et al., 2002; Guilluy et al., 2010; Marrero et al., 1995; Sayeski et al., 2000).
Unlike JAK-STAT signaling associated with cytokine receptors, when AT1R is unbound by
angiotensin II, JAK2 is dissociated from AT1R. Non-activated JAK2 is found in the cytosol and
only upon angiotensin II binding of AT1R does JAK2 autophosphorylation (activation) occur.
Upon activation, phosphorylated JAK2 translocates to and physically associates with AT1R. The
presence of amino acids 319–322 (YIPP) within the carboxyl terminus of AT1R is necessary for the
physical interaction of phosphorylated JAK2 and the AT1R, as mutation of the YIPP sequence was
associated with loss of JAK2-AT1R interaction (Ali et al., 1997). Thus, binding of JAK2 to AT1R
appears to be ligand-dependent, unlike cytokine receptors where JAK2 is constitutively bound,
JAK-STAT Signaling in Cardiovascular Disease 113

marking a major differentiation between JAK-STAT activation between the AT1R and cytokine
receptors (Ali et al., 1998; Sayeski et al., 1999, 2001).
Once the JAK2-AT1R interaction occurs, the latent transcription factor STAT1 is recruited to
the JAK2-AT1R complex (Ali et al., 2000). Although the cytoplasmic carboxyl terminus of AT1R
contains a number of tyrosines, mutations of each of these tyrosines were not found to affect
phosphorylation of STAT1 (Ali et al., 2000). STAT1 is recruited to JAK2 and phosphorylation of
STAT1 at the JAK2-AT1 receptor complex (Sayeski et al., 2001). Moreover, STAT1 phosphoryla­
tion was found to be JAK2-dependent (Ali et al., 2000). Phosphorylation of STAT1 in response to
angiotensin II could be reduced in the presence of the JAK inhibitor, AG490, providing
pharmacological evidence that JAK2, and not the AT1R, per se, mediates phosphorylation of
STAT1 (Ali et al., 1998). Thus, JAK2 is said to act as both a STAT1 kinase and as a molecular
bridge linking STAT1 to the AT1R (Figure 7.3).
As JAK2 deficiency is lethal, insight into the physiological role of JAK2 in hypertension was
initially obtained utilizing pharmacological approaches. Treatment of angiotensin II-infused mice
with the JAK inhibitor, AG490, prevents the development of hypertension and vascular injury
(Neubauer et al., 1998; Guilluy et al., 2010). These studies were complimentary to the in vitro cell
culture studies and serve to implicate an in vivo role for JAK–dependent signaling in hypertension.
To circumvent the limitation associated with the lethality of global JAK2 deficiency, smooth
muscle-specific JAK2 deficient mice were developed (Krempler et al., 2004). Smooth muscle
JAK2-deficiency is associated with reductions in angiotensin II-induced hypertension and
increases in vascular superoxide, a mediator of endothelial dysfunction and vascular hypertrophy

FIGURE 7.3 Angiotensin II-mediated activation of JAK-STAT signaling in vascular smooth muscle. Unlike cytokine
receptors, when the AT1 receptor is unbound, JAK2 is inactive and dissociated from the AT1 receptor. Ang II binding
of the AT1 receptor produces conformational changes in the receptor that results in autophosphorylation and
activation of JAK2. Phosphorylated JAK2 is the recruited to the Ang II bound AT1 receptor. Binding of JAK2 to
the AT1 receptor is dependent upon YRFRR and YIPP consensus sequences in JAK2 and AT1 receptor, respectively.
Mutation of either YRFRR or YIPP consensus sequences results in the inhibition of STAT1 translocation to the AT1
receptor and loss of STAT1-mediated gene transcription. Once JAK2 binds AT1 receptor, STAT1 is recruited to the
JAK2/AT1R complex, where it is phosphorylated before dimerizing and translocating to the nucleus to increase or
decrease transcription of STAT1 target genes. Ang II, angiotensin II; AT1R, angiotensin II type 1 receptor; JAK,
Janus-associated Kinase; STAT, Signal Transducer and Activator of Transcription.
114 JAK-STAT Signaling in Diseases

(Kirabo et al., 2011a). These studies highlight the contribution of JAK2 expression and activation
within the vasculature to promote hypertension-related end-organ injury.
Arhgef1 is a RhoA-specific guanine exchange factor linked to enhanced rho kinase activity in
hypertension. Activated Arhgef1 promotes activation of RhoA and downstream activations of
Rho kinase and enhanced vasoconstriction associated with hypertension. Interestingly, JAK2
activation in response to angiotensin II is associated with phosphorylation and activation of
Arhgef1, but not Arhgef11 or Arhgef12 (Guilluy et al., 2010). JAK inhibition with AG490, JAK2
siRNA, or a dominant negative JAK2 is associated with reduced JAK2 signaling and reduced
angiotensin II-induced phosphorylation of Arhgef1 in cultured smooth muscle (Guilluy et al.,
2010). Moreover, smooth muscle-specific Arhgef1 deficiency was associated with loss of Rho
kinase activity as well as loss of the pressor response to angiotensin II (Guilluy et al., 2010).
Pharmacological inhibition of STAT3 signaling has been demonstrated to limit hypertension
and endothelial dysfunction (Johnson et al., 2013). For example, endothelial dysfunction pro­
duced by angiotensin II was substantially reduced by S3I-201 and STATTIC, two small molecule
inhibitors of STAT3 with slightly different mechanisms of action (Johnson et al., 2013). S3I-201
has high affinity for the SH2 domain of STAT3, but little affinity for STAT1 or STAT5 SH2
domains, thus selectively preventing STAT3 phosphorylation (Siddiquee et al., 2007). STATTIC
directly inhibits STAT3 dimerization thereby directly inhibiting STAT3-mediated gene transcription
(Schust et al., 2006). Interestingly, S3I-201 also prevented angiotensin II-induced increases in
NADPH oxidase-derived superoxide (Johnson et al., 2013).
Like JAK2-deficient mice, homozygous STAT3 deficiency is associated with a lethal phenotype
(Takeda et al., 1997). Thus, genetic insight into the role of STAT3 in hypertension has been
through the use of STAT3-mutant and heterozygous STAT3-deficient mice (Johnson et al., 2013;
Zouein et al., 2013a). Mutant SA/SA mice have reduced, but not complete absence of STAT3
protein or STAT3 activity (Shen et al., 2004). Basal blood pressure is similar in SA/SA mice as
compared to wild-type mice, suggesting that phosphorylation of STAT3 at Ser727 does not
contribute to the regulation of blood pressure (Zouein et al., 2013a). Similarly, the development
of hypertension in response to angiotensin II is not blunted in SA/SA mice (Zouein et al., 2013a).
Despite the general lack of effect of SA/SA on blood pressure, SA/SA appeared to be associated
with a greater degree of myocardial fibrosis and myocyte cell death. These data are supportive of
the concept that STAT3 plays more of a protective role in the heart.
As homozygous SOCS3 deficiency is also lethal, information regarding the role of SOCS3
deficiency has been garnered from studies involving pharmacological inhibition of SOCS3 or
alternatively through the use of mice that are heterozygous deficient for SOCS3 (Li et al., 2016b).
Baseline blood pressure under normal conditions is not altered by loss of a single SOCS3 gene (Li
et al., 2016b). Surprisingly, heterozygous SOCS3 deficiency has no effect on the hypertension
produced by angiotensin II infusion. However, relaxation to acetylcholine in the carotid artery was
impaired in angiotensin II-, but not vehicle-, infused SOCS3 -deficient mice (Li et al., 2016b). These
findings suggest that loss of SOCS3 results in lower inhibition of SOCS3 on STAT3 signaling as
revealed by enhanced vascular effects of angiotensin II in the presence of a single SOCS3 gene. The
lack of effect of heterozygous SOCS3 deficiency on angiotensin-induced hypertension suggests that
the protective effects of SOCS3 on STAT3 expression and activity are complex as reflected by the
observed differences within the vascular wall and on arterial pressure (Li et al., 2016b).

7.8 JAK-STAT in Stroke


Vascular disease, such as carotid artery disease, is a major risk factor for ischemic stroke (Boehme
et al., 2017; Gupta et al., 2013). Although much is known regarding the role of JAK-STAT in
heart failure and hypertension, as discussed above, the role of JAK-STAT in stroke and functional
recovery of neurons post-stroke is limited. Initial studies have found that STAT1 is activated in
response to reactive oxygen species in neurons following ischemia, and deficiency of STAT1 was
JAK-STAT Signaling in Cardiovascular Disease 115

associated with reductions in cerebral infarct size and TUNEL positive staining and improved
neurological scores (Takagi et al., 2002). Such findings suggest that STAT1 expression in neurons
plays a functional role in cerebral injury following ischemia.
In contrast, and somewhat analogous to findings in the heart, STAT3 appears to confer
protection against cerebral injury following ischemic stroke in several different animal models
(Hu et al., 2017; Jung et al., 2009; Liang et al., 2016; Suzuki et al., 2001; Tang et al., 2018; Wang
et al., 2017). For example, middle cerebral artery occlusion is associated with activation of JAK2
and STAT3, which was associated with neuroprotection and angiogenesis (Tang et al., 2018;
Wang et al., 2017). The neuroprotective effects of JAK2 and STAT3 were associated with
increases in epidermal growth factor receptor signaling, which could be inhibited with the JAK
inhibitor, AG-490 (Tang et al., 2018). Neuronal cell death as a result of ischemia is associated with
loss of STAT3 phosphorylation and neuroprotection due in part to loss of manganese superoxide
dismutase, a STAT3 target gene (Jung et al., 2009).
Ischemic stroke produced by middle cerebral artery occlusion is associated with a significant
increase in STAT3 phosphorylation within 6 hours of ischemia in cortical neurons, but not in
astrocytes or microglia (Wen et al., 2001). More recently, it has been suggested that the protective
and detrimental effects of JAK-STAT activation in stroke are most likely related of the temporal
and spatial activation of the various STAT isoforms. For example, the degree of cerebral injury
produced by middle cerebral artery occlusion may actually by reflective of a balance of increased
STAT3 phosphorylation and reduced phosphorylation of STAT6 (Jang et al., 2014). Thus, it may
be possible to pharmacologically manipulate the balance and sequence of STAT activation to
limit the clinical impact associated with stroke. Future studies directed toward understanding the
role of JAK-STAT in stroke are clearly warranted.

7.9 Conclusion
Based on the discussion provided here, it is readily apparent that JAK-STAT plays a critical role in
a number of cardiovascular diseases. While cytokines and cytokine receptors play key roles in
activation of JAK-STAT signaling, growth factors as well as G protein-coupled receptors, such as
angiotensin II and AT1R are also capable of activating JAK-STAT, albeit through slightly different
mechanisms of JAK-STAT recruitment and activation. JAK-STAT is associated with cardiac and
vascular fibrosis and dysfunction. Similarly, JAK-STAT appears to contribute to increases in arterial
blood pressure, which also can have marked effects on cardiovascular structure and function. While
pharmacological inhibitors provided much insight into the functional effects of JAK-STAT signaling,
more selective inhibitors of JAK-STAT are being developed and tested clinically. Similarly, genetically
modified, cell-specific animals have begun to shed some insight into JAK-STAT in a number of
cardiovascular disease models. Finally, while JAK-STAT gain- and loss-of-function polymorphisms
have provided some insight into the role of JAK-STAT in immunity and cancer, to date, there is little
information regarding the effects of JAK-STAT mutations on cardiovascular disease outcomes and is
thus an important and attractive area for additional exploration.

REFERENCES
Ai, D., Jiang, H., Westerterp, M. et al. (2014). Disruption of mammalian target of rapamycin
complex 1 in macrophages decreases chemokine gene expression and atherosclerosis. Circ Res
114:1576–1584.
Ali, M.S., Sayeski, P.P., Bernstein, K.E. (2000). JAK2 acts as both a STAT1 kinase and as a molecular
bridge linking STAT1 to the angiotensin II AT1 receptor. J Biol Chem 275:15586–15593.
Ali, M.S., Sayeski, PP, Dirksen, L.B., Hayzer, D.J., Marrero, M.B., Bernstein, K.E. (1997). Dependence
on the motif YIPP for the physical association of JAK2 kinase with the intracellular carboxy tail of
the angiotensin II AT1 receptor. J Biol Chem 272:23382–23388.
116 JAK-STAT Signaling in Diseases

Ali, M.S., Sayeski, P.P., Safavi, A., Lyles, M., Bernstein, K.E. (1998). Janus kinase 2 (JAK2) must be
catalytically active to associate with the AT1 receptor in response to angiotensin II. Biochem
Biophys Res Comm 249:672–677.
Alturkmani, H.J., Zgheib, C., Zouein, F.A., Alshaaer, N.E., Kurdi, M., Booz, G.W. (2012). Selenate
enhances STAT3 transcriptional activity in endothelial cells: differential actions of selenate and
selenite on LIF cytokine signaling and cell viability. J Inorg Biochem 109:9–15.
Baird, A.M., Thomis, D.C., Berg, L.J. (1998). T cell development and activation in JAK3-deficient mice.
J Leukoc Biol 63:669–677.
Bhat, G.J., Baker, K.M. (1997). Angiotensin II stimulates rapid serine phosphorylation of transcription
factor Stat3. Mol Cell Biochem 170:171–176.
Boehme, A.K., Esenwa, C., Elkind, M.S. (2017). Stroke risk factors, genetics, and prevention. Circ Res
120:472–495.
Bolli, R., Stein, A.B., Guo, Y. et al. (2011). A murine model of inducible, cardiac-specific deletion of
STAT3: its use to determine the role of STAT3 in the upregulation of cardioprotective proteins by
ischemic preconditioning. J Mol Cell Cardiol 50:589–597.
Booz, G.W. (2005). Putting the brakes on cardiac hypertrophy: exploiting the NO-cGMP
counter-regulatory system. Hypertension 45:341–346.
Booz, G.W., Day, J.N., Baker, K.M. (2002). Interplay between the cardiac renin angiotensin system and
JAK–STAT signaling: role in cardiac hypertrophy, ischemia/reperfusion dysfunction, and heart
failure. J Mol Cell Cardiol 34:1443–1453.
Buford, T.W. (2016). Hypertension and aging. Ageing Res Rev 26:96–111.
Chae, C.U., Lee, R.T., Rifai, N., Ridker, P.M. (2001). Blood pressure and inflammation in apparently
healthy men. Hypertension 38:399–403.
Chang, W.T., Cheng, J.T., Chen, Z.C. (2016). Telmisartan improves cardiac fibrosis in diabetes through
peroxisome proliferator activated receptor delta (PPAR delta): from bedside to bench. Cardiovasc
Diabetol 15:113.
Chen, D., Liu, J., Rui, B. et al. (2014). GSTpi protects against angiotensin II-induced proliferation and
migration of vascular smooth muscle cells by preventing signal transducer and activator of
transcription 3 activation. Biochim Biophys Acta 1843:454–463.
Chen, F., Chen, D., Zhang, Y. et al. (2017). Interleukin-6 deficiency attenuates angiotensin II-induced
cardiac pathogenesis with increased myocyte hypertrophy. Biochem Biophys Res Comm 494:534–541.
Chung, C.D., Liao, J., Liu, B. et al. (1997). Specific inhibition of Stat3 signal transduction by PIAS3.
Science 278:1803–1805.
Darnell, J.E. Jr., Kerr, I.M., Stark, G.R. (1994). JAK–STAT pathways and transcriptional activation in
response to IFNs and other extracellular signaling proteins. Science 264:1415–1421.
Das, A., Salloum, F.N., Filippone, S.M. et al. (2015). Inhibition of mammalian target of rapamycin protects
against reperfusion injury in diabetic heart through STAT3 signaling. Basic Res Cardiol 110:31.
Dasgupta, M., Unal, H., Willard, B., Yang, J., Karnik, S.S., Stark, G.R. (2014). Critical role for lysine 685
in gene expression mediated by transcription factor unphosphorylated STAT3. J Biol Chem
289:30763–30771.
Didion, S.P. (2017). Cellular and oxidative mechanisms associated with interleukin-6 signaling in the
vasculature. Int J Mol Sci 18:e2563.
Didion, SP, Faraci, F.M. (2003). Angiotensin II produces superoxide-mediated impairment of endothelial
function in cerebral arterioles. Stroke 34:2038–2042.
Didion, S.P., Kinzenbaw, D.A., Faraci, F.M. (2005). Critical role for CuZn-superoxide dismutase in
preventing angiotensin II-induced endothelial dysfunction. Hypertension 46:1147–1153.
Didion, S.P., Kinzenbaw, D.A., Schrader, L.I., Chu, Y., Faraci, F.M. (2009). Endogenous interleukin-10
inhibits angiotensin II-induced vascular dysfunction. Hypertension 54:619–624.
Didion, S.P., Sigmund, C.D., Faraci, F.M. (2000). Impaired endothelial function in transgenic mice
expressing both human renin and human angiotensinogen. Stroke 31:760–764.
Drazner, M.H. (2011). The progression of hypertensive heart disease. Circulation 123:327–334.
Drenger, B., Ostrovsky, I.A., Barak, M., Nechemia-Arbely, Y., Ziv, E., Axelrod, J.H. (2011). Diabetes
blockade of sevoflurane postconditioning is not restored by insulin in the rat heart: phosphorylated
signal transducer and activator of transcription 3- and phosphatidylinositol 3-kinase-mediated
inhibition. Anesthesiology 114:1364–1372.
JAK-STAT Signaling in Cardiovascular Disease 117

Durbin, J.E., Hackenmiller, R., Simon, M.C., Levy, D.E. (1996). Targeted disruption of the mouse Stat1
gene results in compromised innate immunity to viral disease. Cell 84:443–450.
Dutzmann, J., Daniel, J.M., Bauersachs, J., Hilfiker-Kleiner, D., Seding, D.G. (2015). Emerging transla­
tional approaches to target STAT3 signaling and its impact on vascular disease. Cardiovasc Res
106:365–374.
Enomoto, D., Obana, M., Miyawaki, A., Maeda, M., Nakayama, H., Fujio, Y. (2015). Cardiac-specific
ablation of the STAT3 gene in the subacute phase of myocardial infarction exacerbated cardiac
remodeling. Am J Physiol Heart Circ Physiol 309:H471–480.
Feldman, G.M., Rosenthal, L.A., Liu, X. et al. (1997). STAT5A-deficient mice demonstrate a defect in
granulocyte-macrophage colony-stimulating factor-induced proliferation and gene expression.
Blood 90:1768–1776.
Feuerborn, R., Becker, S., Poti, F. et al. (2017). High density lipoprotein (HDL)-associated sphingosine
1-phosphate (S1P) inhibits macrophage apoptosis by stimulating STAT3 activity and survivin
expression. Atherosclerosis 257:29–37.
Fiaschi, T., Magherini, F., Gamberi, T. et al. (2014). Hyperglycemia and angiotensin II cooperate to
enhance collagen I deposition by cardiac fibroblasts through a ROS-STAT3-dependent mechanism.
Biochim Biophys Acta 1843:2603–2610.
Frank, G.D., Saito, S., Motley, E.D. et al. (2002). Requirement of Ca(2+) and PKCdelta for Janus kinase
2 activation by angiotensin II: involvement of PYK2. Mol Endocrinol 16:367–377.
Frenzel, K., Wallace, T.A., McDoom, I. et al. (2006). A functional JAK2 tyrosine kinase domain is
essential for mouse development. Exp Cell Res 312:2735–2744.
Fulop, T., Larbi, A., Douziech, N., Levesque, I., Varin, A., Herbein, G. (2006). Cytokine receptor
signalling and aging. Mech Ageing Dev 127:526–537.
Gan, X.T., Rajapurohitam, V., Xue, J. et al. (2015). Myocardial hypertrophic remodeling and impaired
left ventricular function in mice with a cardiac-specific deletion of Janus kinase 2. Am J Pathol
185:3202–3210.
Gomolak, J.R., Didion, S.P. (2014). Angiotensin II-induced endothelial dysfunction is temporally linked

with increases in interleukin-6 and vascular macrophage accumulation. Front Physiol 5:396.

Guilluy, C., Bregeon, J., Toumaniantz, G. et al. (2010). The rho exchange factor Arhgef1 mediates the

effects of angiotensin on vascular tone and blood pressure. Nature Med 16:183–190.
Gupta, A., Baradaran, H., Schweitzer, A.D. et al. (2013). Carotid plaque MRI and stroke risk:
a systematic review and meta-analysis. Stroke 44:3071–3077.
Guschin, D, Rogers, N, Briscoe, J et al. (1995). A major role for the protein tyrosine kinase JAK1 in the

JAK/STAT signal transduction pathway in response to interleukin-6. EMBO J 14:1421–1429.

Harvey, A., Montezano, A.C., Lopes, R.A., Rios, F., Touyz, R.M. (2016). Vascular fibrosis in aging and

hypertension: molecular mechanisms and clinical implications. Can J Cardiol 32:659–668.


Hilfiker-Kleiner, D., Hilfiker, A., Fuchs, M. et al. (2004). Signal transducer and activator of transcription
3 is required for myocardial capillary growth, control of interstitial matrix deposition, and heart
protection from ischemic injury. Circ Res 95:187–195.
Hilfiker-Kleiner, D., Shukla, P., Klein, G. et al. (2010). Continuous glycoprotein-130-mediated signal
transducer and activator of transcription-3 activation promotes inflammation, left ventricular
rupture, and adverse outcome in subacute myocardial infarction. Circulation 122:145–155.
Hu, G.Q., Du, X., Li, Y.J., Gao, X.Q., Chen, B.Q., Yu, L. (2017). Inhibition of cerebral ischemia/
reperfusion injury-induced apoptosis: nicotiflorin and JAK2/STAT3 pathway. Neural Regener Res
12:96–102.
Hubbard, S.R. (2017). Mechanistic insights into regulation of JAK2 tyrosine kinase. Front Endocrinol
(Lausanne) 8:361.
Hussain, M., Awan, F.R. (2018). Hypertension regulating angiotensin peptides in the pathobiology of
cardiovascular disease. Clin Exp Hypertens 40:344–352.
Jacoby, J.J., Kalinowski, A., Liu, M.G. et al. (2003). Cardiomyocyte-restricted knockout of STAT3 results
in higher sensitivity to inflammation, cardiac fibrosis, and heart failure with advanced age. Proc
Natl Acad Sci USA 100:12929–12934.
Jang, S.S., Choi, J.H., Im, D.S. et al. (2014). The phosphorylation of STAT6 during ischemic reperfusion
in rat cerebral cortex. Neuroreport 25:18–22.
118 JAK-STAT Signaling in Diseases

Johnson, A.W., Kinzenbaw, D.A., Modrick, M.L., Faraci, F.M. (2013). Small molecule inhibitors of
STAT3 protect against angiotensin II-induced vascular dysfunction and hypertension. Hypertension
61:437–442.
Jung, J.E., Kim, G.S., Narasimhan, P., Song, Y.S., Chan, P.H. (2009). Regulation of Mn-superoxide
dismutase activity and neuroprotection by STAT3 in mice after cerebral ischemia. J Neurosci
29:7003–7014.
Jusufovic, M., Sandset, E.C., Skagen, K., Skjelland, M. (2016). Blood pressure lowering treatment in
patients with carotid artery stenosis. Curr Hypertens Rev 12:148–155.
Kakutani, Y., Shioi, A., Shoji, T. et al. (2015). Oncostatin M promotes osteoblastic differentiation of human
vascular smooth muscle cells through JAK3-STAT3 pathway. J Cell Biochem 116:1325–1333.
Kano, A., Wolfgang, M.J., Gao, Q. et al. (2003). Endothelial cells require STAT3 for protection against
endotoxin-induced inflammation. J Exp Med 198:1517–1525.
Kaplan, M.H., Schindler, U., Smiley, S.T., Grusby, M.J. (1996a). Stat6 is required for mediating responses
to IL-4 and for development of Th2 cells. Immunity 4:313–319.
Kaplan, M.H., Sun, Y.L., Hoey, T., Grusby, M.J. (1996b). Impaired IL-12 responses and enhanced
development of Th2 cells in Stat4-deficient mice. Nature 382:174–177.
Kaptein, A., Paillard, V., Saunders, M. (1996). Dominant negative STAT3 mutant inhibits interleukin-
6-induced JAK–STAT signal transduction. J Biol Chem 271:5961–5964.
Kawai, T., Forrester, S.J., O’Brien, S., Baggett, A., Rizzo, V., Eguchi, S. (2017). AT1 receptor signaling
pathways in the cardiovascular system. Pharmacol Res 125:4–13.
Kile, B.T., Alexander, W.S. (2001). The suppressors of cytokine signalling (SOCS). Cell Mol Life Sci
58:1627–1635.
Kirabo, A., Kearns, P.N., Jarajapu, Y.P. et al. (2011a). Vascular smooth muscle JAK2 mediates angiotensin
II-induced hypertension via increased levels of reactive oxygen species. Cardiovasc Res 91:171–179.
Kirabo, A., Oh, S.P., Kasahara, H., Wagner, K.U., Sayeski, P.P. (2011b). Vascular smooth muscle JAK2
deletion prevents angiotensin II-mediated neointima formation following injury in mice. J Mol Cell
Cardiol 50:1026–1034.
Kirchmer, M.N., Franco, A., Albasanz-Puig, A. et al. (2014). Modulation of vascular smooth muscle cell
phenotype by STAT-1 and STAT-3. Atherosclerosis 234:169–175.
Krebs, D.L., Hilton, D.J. (2001). SOCS proteins: negative regulators of cytokine signaling. Stem Cells
19:378–387.
Krempler, A., Qi, Y., Triplett, A.A., Zhu, J., Rui, H., Wagner, K.U. (2004). Generation of a conditional
knockout allele for the Janus kinase 2 (JAK2) gene in mice. Genesis 40:52–57.
Kunisada, K., Negoro, S., Tone, E. et al. (2000). Signal transducer and activator of transcription 3 in the
heart transduces not only a hypertrophic signal but a protective signal against doxorubicin-induced
cardiomyopathy. Proc Natl Acad Sci USA 97:315–319.
Kurdi, M., Booz, G.W. (2007). Can the protective actions of JAK–STAT in the heart be exploited
therapeutically? Parsing the regulation of interleukin-6-type cytokine signaling. J Cardiovasc
Pharmacol 50:126–141.
Lai, N.C., Gao, M.H., Tang, E. et al. (2012). Pressure overload-induced cardiac remodeling and
dysfunction in the absence of interleukin 6 in mice. Lab Invest 92:1518–1526.
Li, H., Liu, W., Xie, J. (2017). Circulating interleukin-6 levels and cardiovascular and all-cause mortality
in the elderly population: a meta-analysis. Arch Gerontol Geriatr 73:257–262.
Li, H., Yao, W., Liu, Z. et al. (2016a). Hyperglycemia abrogates ischemic postconditioning cardioprotec­
tion by impairing AdipoR1/Caveolin-3/STAT3 signaling in diabetic rats. Diabetes 65:942–955.
Li, S., Geng, Q., Chen, H. et al. (2018). The potential inhibitory effects of miR19b on vulnerable plaque
formation via the suppression of STAT3 transcriptional activity. Int J Mol Med 41:859–867.
Li, Y., Kinzenbaw, D.A., Modrick, M.L., Pewe, L.L., Faraci, F.M. (2016b). Context-dependent effects of
SOCS3 in angiotensin II-induced vascular dysfunction and hypertension in mice: mechanisms and
role of bone marrow-derived cells. Am J Physiol Heart Circ Physiol 311:H146–56.
Liang, Z., Wu, G., Fan, C. et al. (2016). The emerging role of signal transducer and activator of
transcription 3 in cerebral ischemic and hemorrhagic stroke. Prog Neurobiol 137:1–16.
Liu, S., Liu, L.X., Zhang, Y.L. et al. (2018). Cardiac ablation of SOCS3 aggravates DOCA-salt-induced
hypertrophic remodeling by activation of gp130-dependent signaling in mice. Cell Physiol Biochem
47:140–150.
JAK-STAT Signaling in Cardiovascular Disease 119

Liu, X., Robinson, G.W., Wagner, K.U., Garrett, L., Wynshaw-Boris, A., Hennighausen, L. (1997). Stat5a
is mandatory for adult mammary gland development and lactogenesis. Genes Dev 11:179–186.
Lo Coco, D., Lopez, G., Corrao, S. (2016). Cognitive impairment and stroke in elderly patients. Vasc
Health Risk Manag 12:105–116.
Lo, S.H., Hsu, C.T., Niu, H.S., Niu, C.S., Cheng, J.T., Chen, Z.C. (2017). Cryptotanshinone inhibits
STAT3 signaling to alleviate cardiac fibrosis in type 1-like diabetic rats. Phytother Res 31:638–646.
Longo, D.M., Louie, B., Putta, S. et al. (2012). Single-cell network profiling of peripheral blood mono­
nuclear cells from healthy donors reveals age- and race-associated differences in immune signaling
pathway activation. J Immunol 188:1717–1725.
Madamanchi, N.R., Li, S., Patterson, C., Runge, M.S. (2001). Thrombin regulates vascular smooth muscle
cell growth and heat shock proteins via the JAK–STAT pathway. J Biol Chem 276:18915–18924.
Maggio, M., Guralnik, J.M., Longo, D.L., Ferrucci, L. (2006). Interleukin-6 in aging and chronic disease:
a magnificent pathway. J Gerontol A Biol Sci Med Sci 61:575–584.
Marrero, M.B. (2005). Introduction to JAK/STAT signaling and the vasculature. Vascul Pharmacol
43:307–309.
Marrero, M.B., Paxton, W.G., Schieffer, B., Ling, B.N., Bernstein, K.E. (1996). Angiotensin II signalling
events mediated by tyrosine phosphorylation. Cell Signal 8:21–26.
Marrero, M.B., Schieffer, B., Li, B., Sun, J., Harp, J.B., Ling, B.N. (1997). Role of Janus kinase/signal
transducer and activator of transcription and mitogen-activated protein kinase cascades in angio­
tensin II- and platelet-derived growth factor-induced vascular smooth muscle cell proliferation.
J Biol Chem 272:24684–24690.
Marrero, M.B., Schieffer, B., Paxton, W.G. et al. (1995). Direct stimulation of JAK/STAT pathway by the
angiotensin II AT1 receptor. Nature 375:247–250.
McCormick, J., Suleman, N., Scarabelli, T.M., Knight, R.A., Latchman, D.S., Stephanou, A. (2012).
STAT1 deficiency in the heart protects against myocardial infarction by enhancing autophagy.
J Cell Mol Med 16:386–393.
McWhinney, C.D., Hunt, R.A., Conrad, K.M., Dostal, D.E., Baker, K.M. (1997). The type I angiotensin
II receptor couples to Stat1 and Stat3 activation through JAK2 kinase in neonatal rat cardiac
myocytes. J Mol Cell Cardiol 29:2513–2524.
Mehta, P.K., Griendling, K.K. (2007). Angiotensin II cell signaling: physiological and pathological effects
in the cardiovascular system. Am J Physiol Cell Physiol 292:C82–897.
Meraz, M.A., White, J.M., Sheehan, K.C. et al. (1996). Targeted disruption of the Stat1 gene in mice
reveals unexpected physiologic specificity in the JAK–STAT signaling pathway. Cell 84:431–442.
Nagai, M., Hoshide, S., Kario, K. (2010). Hypertension and dementia. Am J Hypertens 23:116–124.
Nagata, T., Yasukawa, H., Kyogoku, S. et al. (2015). Cardiac-specific SOCS3 deletion prevents in vivo
myocardial ischemia reperfusion injury through sustained activation of cardioprotective signaling
molecules. PLoS One 10:e0127942.
Nakagawa, K., and Nakashima, Y. (2018). Pathologic intimal thickening in human atherosclerosis is
formed by extracellular accumulation of plasma-derived lipids and dispersion of intimal smooth
muscle cells. Atherosclerosis 274:235–242.
Neubauer, H., Cumano, A., Müller, M., Wu, H., Huffstadt, U., Pfeffer, K. (1998). JAK2 deficiency defines
an essential developmental checkpoint in definitive hematopoiesis. Cell 93:397–409.
Nosaka, T., van Deursen, J.M., Tripp, R.A. et al. (1995). Defective lymphoid development in mice lacking
JAK3. Science 270:800–802.
O’Brown, Z.K., Van Nostrand, E.L., Higgins, J.P., Kim, S.K. (2015). The inflammatory transcription
factors NFκB, STAT1 and STAT3 drive age-associated transcriptional changes in the human
kidney. PLoS Genet 11:e1005734.
Oba, T., Yasukawa, H., Hoshijima, M. et al. (2012). Cardiac-specific deletion of SOCS-3 prevents
development of left ventricular remodeling after acute myocardial infarction. J Am Coll Cardiol
59:838–852.
Okuda, K., Foster, R., Griffin, J.D. (1999). Signaling domains of the beta c chain of the GM-CSF/IL-3/
IL-5 receptor. Ann N Y Acad Sci 872:305–312.
Oliverio, M.I., Best, C.F., Kim, H.S., Arendshorst, W.J., Smithies, O., Coffman, T.M. (1997). Angiotensin
II responses in AT1A receptor-deficient mice: a role for AT1B receptors in blood pressure
regulation. Am J Physiol 272:F515–20.
120 JAK-STAT Signaling in Diseases

Owais, K., Huang, T., Mahmood, F. et al. (2015). Cardiopulmonary bypass decreases activation of the
signal transducer and activator of transcription 3 (STAT3) pathway in diabetic human
myocardium. Ann Thorac Surg 100:1636–1645.
Pan, J., Fukuda, K., Saito, M. et al. (1999). Mechanical stretch activates the JAK/STAT pathway in rat
cardiomyocytes. Circ Res 84:1127–1136.
Parganas, E., Wang, D., Stravopodis, D. et al. (1998). JAK2 is essential for signaling through a variety of
cytokine receptors. Cell 93:385–395.
Park, C., Li, S., Cha, E., Schindler, C. (2000). Immune response in Stat2 knockout mice. Immunity
13:795–804.
Park, S.O., Wamsley, H.L., Bae, K. et al. (2013). Conditional deletion of JAK2 reveals an essential role in
hematopoiesis throughout mouse ontogeny: implications for JAK2 inhibition in humans. PLoS
One 8:e59675.
Pasquin, S., Laplante, V., Kouadri, S. et al. (2018). Cardiotrophin-like cytokine increases
macrophage-foam cell transition. J Immunol 201:2462–2471.
Patel, P., Ordunez, P., DiPette, D. et al. (2016). Improved blood pressure control to reduce cardiovascular
disease morbidity and mortality: The standardized hypertension treatment and prevention project.
J Clin Hypertens 18:1284–1294.
Paulus, W.J., Tschope, C. (2013). A novel paradigm for heart failure with preserved ejection fraction:
comorbidities drive myocardial dysfunction and remodeling through coronary microvascular
endothelial inflammation. J Am Coll Cardiol 62:263–271.
Pugach, E.K., Richmond, P.A., Azofeifa, J.G., Dowell, R.D., Leinwand, L.A. (2015). Prolonged Cre
expression driven by the alpha-myosin heavy chain promoter can be cardiotoxic. J Mol Cell Cardiol
86:54–61.
Recinos, A. III, LeJeune, W.S., Sun, H. et al. (2007). Angiotensin II induces IL-6 expression and the
JAK–STAT3 pathway in aortic adventitia of LDL receptor-deficient mice. Atherosclerosis
194:125–133.
Rodig, S.J., Meraz, M.A., White, J.M. et al. (1998). Disruption of the JAK1 gene demonstrates obligatory
and nonredundant roles of the JAKs in cytokine-induced biologic responses. Cell 93:373–383.
Rosendorff, C., Lackland, D.T., Allison, M. et al. (2015). Treatment of hypertension in patients with
coronary artery disease: a scientific statement from the American heart association, American
college of cardiology, and American society of hypertension. Circulation 131:e435–70.
Ryan, M.J., Didion, S.P., Mathur, S., Faraci, F.M., Sigmund, C.D. (2004). Angiotensin II-induced

vascular dysfunction is mediated by the AT1A receptor in mice. Hypertension 43:1074–1079.

Sakamoto, K., Wehde, B.L., Rädler, P.D., Triplett, A.A., Wagner, K.U. (2016). Generation of janus kinase

1 (JAK1) conditional knockout mice. Genesis 54:582–588.


Sayeski, P.P., Ali, M.S., Bernstein, K.E. (2000). The role of Ca2+ mobilization and heterotrimeric
G protein activation in mediating tyrosine phosphorylation signaling patterns in vascular smooth
muscle cells. Mol Cell Biochem 212:91–98.
Sayeski, P.P., Ali, M.S., Frank, S.J., Bernstein, K.E. (2001). The angiotensin II-dependent nuclear
translocation of Stat1 is mediated by the JAK2 protein motif 231YRFRR. J Biol Chem
276:10556–10563.
Sayeski, P.P., Ali, M.S., Safavi, A. et al. (1999). A catalytically active JAK2 is required for the angiotensin
II-dependent activation of fyn. J Biol Chem 274:33131–33142.
Schrader, L.I., Kinzenbaw, D.A., Johnson, A.W., Faraci, F.M., Didion, S.P. (2007). IL-6 deficiency
protects against angiotensin II induced endothelial dysfunction and hypertrophy. Arterioscler
Thromb Vasc Biol 27:2576–2581.
Schust, J., Sperl, B., Hollis, A., Mayer, T.U., Berg, T. (2006). Stattic: a small-molecule inhibitor of STAT3
activation and dimerization. Chem Biol 13:1235–1242.
Shen, Y., Schlessinger, K., Zhu, X. et al. (2004). Essential role of STAT3 in postnatal survival and growth
revealed by mice lacking STAT3 serine 727 phosphorylation. Mol Cell Biol 24:407–419.
Shen-Orr, S.S., Furman, D., Kidd, B.A. et al. (2016). Defective signaling in the JAK–STAT pathway
tracks with chronic inflammation and cardiovascular risk in aging humans. Cell Syst 3:374–84.e4.
Shimoda, K., van Deursen, J., Sangster, M.Y. et al. (1996). Lack of IL-4-induced Th2 response and IgE
class switching in mice with disrupted Stat6 gene. Nature 380:630–633.
Shuai, K. (2006). Regulation of cytokine signaling pathways by PIAS proteins. Cell Res 16:196–202.
JAK-STAT Signaling in Cardiovascular Disease 121

Siddiquee, K., Zhang, S., Guida, W.C. et al. (2007). Selective chemical probe inhibitor of Stat3, identified
through structure-based virtual screening, induces antitumor activity. Proc Natl Acad Sci USA
104:7391–7396.
Smit, L.S., Vanderkuur, J.A., Stimage, A. et al. (1997). Growth hormone-induced tyrosyl phosphorylation
and deoxyribonucleic acid binding activity of Stat5A and Stat5B. Endocrinology 138:3426–3434.
Smith, R.M., Suleman, N., Lacerda, L. et al. (2004). Genetic depletion of cardiac myocyte STAT-3
abolishes classical preconditioning. Cardiovasc Res 63:611–616.
Su, S.A., Yang, D., Wu, Y. et al. (2017). EphrinB2 regulates cardiac fibrosis through modulating the
interaction of Stat3 and TGF-beta/Smad3 signaling. Circ Res 121:617–627.
Sun, J., Zhang, M., Chen, K. et al. (2018). Suppression of TLR4 activation by resveratrol is associated
with STAT3 and Akt inhibition in oxidized low-density lipoprotein-activated platelets. Eur
J Pharmacol 836:1–10.
Suzuki, S., Tanaka, K., Nogawa, S. et al. (2001). Phosphorylation of signal transducer and activator of
transcription-3 (Stat3) after focal ischemia in rats. Exp Neurol 170:63–71.
Szczepanek, K., Chen, Q., Derecka, M. et al. (2011). Mitochondrial-targeted Signal transducer and
activator of transcription 3 (STAT3) protects against ischemia-induced changes in the electron
transport chain and the generation of reactive oxygen species. J Biol Chem 286:29610–29620.
Szczepanek, K., Xu, A., Hu, Y. et al. (2015). Cardioprotective function of mitochondrial-targeted and
transcriptionally inactive STAT3 against ischemia and reperfusion injury. Basic Res Cardiol 110:53.
Sztriha, L.K., Nemeth, D., Sefcsik, T., Vecsei, L. (2009). Carotid stenosis and the cognitive function.
J Neurol Sci 283:36–40.
Takagi, Y., Harada, J., Chiarugi, A., Moskowitz, M.A. (2002). STAT1 is activated in neurons after
ischemia and contributes to ischemic brain injury. J Cereb Blood Flow Metabol 22:1311–1318.
Takeda, K., Noguchi, K., Shi, W. et al. (1997). Targeted disruption of the mouse Stat3 gene leads to early
embryonic lethality. Proc Natl Acad Sci USA 94:3801–3804.
Tang, Y., Tong, X., Li, Y. et al. (2018). JAK2/STAT3 pathway is involved in the protective effects of
epidermal growth factor receptor activation against cerebral ischemia/reperfusion injury in rats.
Neurosci Lett 662:219–226.
Te Riet, L., van Esch, J.H., Roks, A.J., van Den Meiracker, A.H., Danser, A.H. (2015). Hypertension:
renin-angiotensin-aldosterone system alterations. Circ Res 116:960–975.
Thomis, D.C., Gurniak, C.B., Tivol, E., Sharpe, A.H., Berg, L.J. (1995). Defects in B lymphocyte
maturation and T lymphocyte activation in mice lacking JAK3. Science 270:794–797.
Tschope, C., Van Linthout, S. (2014). New insights in (inter)cellular mechanisms by heart failure with
preserved ejection fraction. Curr Heart Fail Rep 11:436–444.
Udy, G.B., Towers, R.P., Snell, R.G. et al. (1997). Requirement of STAT5b for sexual dimorphism of body
growth rates and liver gene expression. Proc Natl Acad Sci USA 94:7239–7244.
Ungureanu, D., Wu, J., Pekkala, T. et al. (2011). The pseudokinase domain of JAK2 is a dual-specificity
protein kinase that negatively regulates cytokine signaling. Nat Struct Mol Biol 18:971–976.
van Keulen, D., Pouwer, M.G., Pasterkamp, G. et al. (2018). Inflammatory cytokine oncostatin M induces
endothelial activation in macro- and microvascular endothelial cells and in APOE*3Leiden.CETP
mice. PLoS One 13:e0204911.
Verma, S.K., Krishnamurthy, P., Barefield, D. et al. (2012). Interleukin-10 treatment attenuates pressure
overload-induced hypertrophic remodeling and improves heart function via signal transducers and
activators of transcription 3-dependent inhibition of nuclear factor-kappaB. Circulation 126:418–429.
Wang, L., Li, J., Li, D. (2015). Losartan reduces myocardial interstitial fibrosis in diabetic cardiomyo­
pathy rats by inhibiting JAK/STAT signaling pathway. Int J Clin Exp Pathol 8:466–473.
Wang, R., Zhang, Y., Xu, L. et al. (2016). Protein inhibitor of activated STAT3 suppresses oxidized
LDL-induced cell responses during atherosclerosis in apolipoprotein E-deficient mice. Sci Rep
6:36790.
Wang, T., Mao, X., Li, H. et al. (2013). N-Acetylcysteine and allopurinol up-regulated the JAK/STAT3
and PI3K/Akt pathways via adiponectin and attenuated myocardial postischemic injury in diabetes.
Free Radic Biol Med 63:291–303.
Wang, X.L., Qiao, C.M., Liu, J.O., Li, C.Y. (2017). Inhibition of the SOCS1-JAK2-STAT3 signaling
pathway confers neuroprotection in rats with ischemic stroke. Cell Physiol Biochem 44:85–98.
Wen, S.W., Wong, C.H. (2018). Aging- and vascular-related pathologies. Microcirculation 30:e12463.
122 JAK-STAT Signaling in Diseases

Wen, T.C., Peng, H., Hata, R., Desaki, J., Sakanaka, M. (2001). Induction of phosphorylated-Stat3
following focal cerebral ischemia in mice. Neurosci Lett 303:153–156.
Wiejak, J., Dunlop, J., Mackay, S.P., Yarwood, S.J. (2013). Flavanoids induce expression of the suppressor
of cytokine signalling 3 (SOCS3) gene and suppress IL-6-activated signal transducer and activator
of transcription 3 (STAT3) activation in vascular endothelial cells. Biochem J 454:283–293.
Wolf-Maier, K., Cooper, R.S., Banegas, J.R. et al. (2003). Hypertension prevalence and blood pressure
levels in 6 European countries, Canada, and the United States. JAMA 289:2363–2369.
Wollert, K.C., Taga, T., Saito, M. et al. (1996). Cardiotrophin-1 activates a distinct form of cardiac muscle
cell hypertrophy. Assembly of sarcomeric units in series VIA gp130/leukemia inhibitory factor
receptor-dependent pathways. J Biol Chem 271:9535–9545.
Xu, D., Qu, C.K. (2008). Protein tyrosine phosphatases in the JAK/STAT pathway. Front Biosci
13:4925–4932.
Xu, M., Tchkonia, T., Ding, H. et al. (2015). JAK inhibition alleviates the cellular senescence-associated
secretory phenotype and frailty in old age. Proc Natl Acad Sci USA 112:E6301–10.
Xue, R., Lei, S., Xia, Z.Y. et al. (2016). Selective inhibition of PTEN preserves ischaemic
post-conditioning cardioprotection in STZ-induced type 1 diabetic rats: role of the PI3K/Akt and
JAK2/STAT3 pathways. Clin Sci (Lond) 130:377–392.
Yajima, T., Murofushi, Y., Zhou, H. et al. (2011). Absence of SOCS3 in the cardiomyocyte increases
mortality in a gp130-dependent manner accompanied by contractile dysfunction and ventricular
arrhythmias. Circulation 124:2690–2701.
Yang, P., Zhang, Y., Pang, J. et al. (2013). Loss of JAK2 impairs endothelial function by attenuating
Raf-1/MEK1/Sp-1 signaling along with altered eNOS activities. Am J Pathol 183:617–625.
Yue, H., Li, W., Desnoyer, R., Karnik, S.S. (2010). Role of nuclear unphosphorylated STAT3 in
angiotensin II type 1 receptor-induced cardiac hypertrophy. Cardiovasc Res 85:90–99.
Zgheib, C., Zouein, F.A., Kurdi, M., Booz, G.W. (2012). Differential STAT3 signaling in the heart: impact
of concurrent signals and oxidative stress. JAKSTAT 1:101–110.
Zhang, W., Qu, X., Chen, B. et al. (2016). Critical Roles of STAT3 in beta-adrenergic functions in the
heart. Circulation 133:48–61.
Zhao, L., Cheng, G., Jin, R. et al. (2016). Deletion of interleukin-6 attenuates pressure overload-induced
left ventricular hypertrophy and dysfunction. Circ Res 118:1918–1929.
Zhao, Y., Yan, L., Peng, L. et al. (2018). Oleoylethanolamide alleviates macrophage formation via
AMPK/PPARalpha/STAT3 pathway. Pharmacol Rep 70:1185–1194.
Zile, M.R., Baicu, C.F., Ikonomidis, J.S. et al. (2015). Myocardial stiffness in patients with heart failure
and a preserved ejection fraction: contributions of collagen and titin. Circulation 131:1247–1259.
Zouein, F.A., Altara, R., Chen, Q. et al. (2015). Pivotal importance of STAT3 in protecting the heart from

acute and chronic stress: new advancement and unresolved issues. Front Cardiovasc Med 2:36.

Zouein, F.A., Kurdi, M., Booz, G.W. (2013b). Dancing rhinos in stilettos: The amazing saga of the

genomic and nongenomic actions of STAT3 in the heart. JAKSTAT 2:e24352.


Zouein, F.A., Zgheib, C., Hamza, S. et al. (2013a). Role of STAT3 in angiotensin II-induced hypertension
and cardiac remodeling revealed by mice lacking STAT3 serine 727 phosphorylation. Hypertens Res
36:496–503.
8
Diabetes and Obesity: Abnormal JAK-STAT
Signaling

Marcia J. Abbott
Department of Health Sciences
Crean College of Health and Behavioral Sciences, Chapman University
Orange, California

8.1 Diabetes and Obesity: Abnormal JAK-STAT Signaling


8.1.1 Obesity
Obesity has become a world-wide epidemic and rates continue to rise exponentially. According
to the Centers for Disease Control (CDC), in 2015–2016 the prevalence of obesity was 39.8%
and affecting roughly 93.3 million adults in the United States. Further, these high rates of
obesity resulted in approximately $315.8 billion in healthcare costs (Biener et al., 2017).
Sustained obesity results in the progression of many diseases and disorders, such as, but not
limited to, diabetes, heart disease, and certain cancers, some of the leading preventable diseases
resulting in premature death. Studies suggest that the rates of obesity will continue to rise in all
populations across all demographics. As such, uncovering treatments for obesity is of grave
importance.
Obesity occurs when dietary energy intake is greater than whole body energy expenditure. The
excess energy is stored as fat, with its primary storage site being the adipose tissue, and induces
expansion of fat mass. There are many known treatments for obesity, such as modulations of food
intake and physical activity. However, maintenance of such interventions remains a major road
block for human health. As a result, the scientific community has attempted to unravel the
molecular players that mediate whole body metabolism and food intake. There have been previous
attempts to develop pharmacological treatments for obesity, but many of these compounds have
resulted in serious side effects and even death. Therefore, the scientific community continues to
examine the molecular pathways that contribute to obesity.

8.1.2 Diabetes
Type 2 diabetes (T2D) is one of many diseases resulting from obesity which poses great determents
to overall health. According to the American Diabetes Association, diabetes is the 7th leading cause
of death, with 30.3 million Americans afflicted by the disease, in the United States in 2015.
Importantly, according to the World Health Organization, 90% of people afflicted with T2D are
obese. Fundamentally, T2D occurs as a result of dysregulation of the pancreatic β-cells to sense and
respond to increased blood glucose levels occurring postprandially. Of the two types of diabetes,
Type 1 (T1D) progresses as a result of “death” of pancreatic islets and is primarily due to
autoimmune disorders. For the purposes here, the focus will remain on the association of obesity
and T2D. However, the end result of T2D is a loss of islet function similar to T1D. The onset and
progression of T2D can result in many other conditions such as neuropathy, cardiovascular disease,

123
124 JAK-STAT Signaling in Diseases

retinopathy, and renal disease. One of the major culprits for the development of these aforementioned
conditions is sustained hyperglycemia. Hyperglycemia is a condition that occurs “naturally” in the
postprandial state and is counterbalanced by insulin secretion from pancreatic β-cells. Insulin action
stimulates glucose uptake into the skeletal muscle (SKM), liver, and adipose tissue while inducing
lipogenesis in these tissues. Under obese conditions, many processes charged with mediating glucose
homeostasis become disrupted. Insulin resistance is one of the first signs of T2D and can occur in
many tissues such as adipose tissue and SKM. Insulin resistance tends to develop with the progression
of obesity and inflammatory processes. Even though it is known that a major population of patients
with T2D is obese, and obesity is a confirmed risk factor for the development of T2D, the exact link
between obesity and T2D has yet to be identified and remains complex and multifactorial.

8.1.3 JAK/STAT Signaling


Historically, the Janus kinase (JAK)/Signal transducer and activator of transcription (STAT)
signaling pathway has been studied in response to cytokine signaling in invertebrates (Schindler,
2002). A common feature of cytokines is their ability to bind to their specific cell surface receptor
inducing a conformational transition of the receptor structure (Schindler, 2002). Subsequent to
receptor binding, numerous intracellular targets become activated, including, but not limited to
the JAK family of kinases. Upon cytokine binding, JAKs translocate and bind to the intracellular
region of the target receptor and become auto-phosphorylated (Darnell, 1997). Following activa­
tion, JAKs exert their stimulatory action by phosphorylating tyrosine residues of their substrate
proteins. Four JAK isoforms have been identified JAKs 1–3 and Tyk2. These kinases initiate their
intracellular responses following cytokine activation of their specific plasma membrane-docked
receptors. All but JAK3, proteins within the JAK family, are ubiquitously expressed with their
highest expression found in hematopoietic cells (Ihle et al., 1995). However, recently JAK3 has
been confirmed to be expressed in other cell types, such as SKM (Krolopp et al., 2016). Further,
JAK3 has been found to require binding of the effector cytokine to the γc chain of the target
receptor (Ihle et al., 1995; Miyazaki et al., 1994).
STAT proteins act as intermediary targets of JAKs and in their inactivated state remain
sequestered in the cytoplasm, requiring translocation to the nucleus to exert their transcriptional
regulation. STATs are ubiquitously expressed and have been shown to be both activators and
negative regulators of gene expression. There are seven distinct STAT isoforms, STAT1, 2, 3, 4,
5A, 5B and 6. STAT transcriptional activity has been linked to many cellular pathways, such as
interferon signaling, anti-inflammatory regulation, and a variety of other immune reactions.
Activation of STATs is a transient process which is regulated by tyrosine phosphatases (Takeda
and Akira, 2000).
Complicating JAK/STAT signaling is the multiple combinations of JAK proteins that have the
ability to interact and activate various STAT proteins. It does not appear that STAT proteins are
regulated by one particular JAK. Therefore, it is not clear if these variant interactions, between
JAK and STAT isoforms, play independent regulatory roles or have redundant signaling path­
ways. In summary, activation of JAK/STATs is initiated by binding of the specific ligand to the
receptor on the surface of the target tissue. Following ligand binding, a JAK isoform will become
activated, via auto-phosphorylation, through its interaction with the receptor. The activated JAK
will then phosphorylate its target STAT, on a tyrosine residue, leading to dimerization of the
STAT protein and its transport to the nucleus. Finally, the activated STAT will bind to DNA to
promote/inhibit transcription of various genes in various tissues (Figure 8.1).

8.2 Overview of JAK/STAT in Metabolic Diseases


Over the past decades, JAK/STAT signaling pathways have been attributed to playing a role in
the development and progression of metabolic diseases, such as obesity and subsequently T2D.
Diabetes and Obesity 125

FIGURE 8.1 Activation of JAK/STAT signaling pathways play roles in mediating metabolic processes in many tissues.

JAK/STATs exert their regulatory control in various cells and tissues throughout the body (Figure
8.1). Of importance to their regulation on metabolism is their action in adipocytes, myocytes,
pancreatic β-cells, and hepatocytes. There is evidence that JAK/STAT signaling acts to reduce or
prevent adipocyte differentiation and proliferation in in vitro and in vivo mouse models. On the
other hand, in some situations, STAT signaling can act to induce adipogenesis and weight gain.
Further, complicating JAK/STAT signaling effects on adipose tissue expansion is the discovery that
JAK/STAT pathways have been confirmed to increase the energy expending activity of brown
adipocytes. In SKM cells, activation of JAK/STAT signaling plays a role in stimulating myogenesis,
thus, increasing lean mass, contributing to reductions in overall body fatness. Additionally, JAK/
STATs are found to improve insulin secretion in pancreatic β-cells. Finally, JAK/STATs have been
linked to regulating hepatocyte glucose metabolism, through mediating gluconeogenic pathways. All
of these aforementioned roles of JAK/STAT signaling are complex and point to the need for
uncovering the molecular intermediaries to fully elucidate their role in metabolic processes.
Recently, genetic manipulation of JAKs and STATs, in mouse models, has brought to light the
contribution of these molecular mediators in regulating whole body metabolic processes. Because
regulation of the JAK/STAT pathway is complex and varies among isoforms and tissues, many mouse
models have been developed. Whole body knockout of JAK3 yields increases in body weight,
hyperglycemia and liver dysfunction (Mishra et al., 2015). Similarly, Tyk2 knockout mice display
increased blood glucose as well as reduced glucose tolerance (Derecka et al., 2012). On the other
hand, ubiquitous deletion of STAT4 increases glucose tolerance, insulin sensitivity, with no effects on
body weight, while the mice were on a high fat diet (Dobrian et al., 2013). Independent of STAT4
deletion, STAT6 knockout mice displayed a contradictory phenotype with reductions in body weight,
attributed to an increase in energy expenditure and glucose tolerance (Ricardo-Gonzalez et al., 2010).
Beyond studies in rodent models, associations between JAK2, body fat and waist circumference have
been observed in humans, suggesting JAK2 as a potential target for the treatment of obesity (Ge et al.,
2008). The onset of development of mouse experimental models has begun to shed light on the role
that JAK/STAT signaling plays in whole body metabolic processes, but the molecular regulators
downstream of STAT transcriptional activity have yet to be fully identified. For example, as
previously mentioned, it is clear that STAT4 acts to maintain glucose tolerance; however, the gene
alterations that orchestrate these effects remain unknown. In this regard, there is much room for
126 JAK-STAT Signaling in Diseases

examination of the genes that are altered due to manipulations of STAT signaling to fully
understand the role of these signaling pathways in metabolic processes. Mechanistic, in vitro
studies to complement the in vivo physiological studies will assist in determining the entire
regulatory picture of JAK/STAT signaling. Unarguably, aberrations in JAK and/or STAT signal­
ing results in varied phenotypes in a whole-body metabolic context.

8.2.1 Adipose Tissue and Obesity


Historically, adipose tissue was thought to act solely as a storage site for excess energy in the form of
triaclyglycerides (TAG) stored in specialized organelles, termed lipid droplets. As such, energy
balance is tightly regulated and adipose tissue plays a central role in maintaining homeostasis.
White adipose tissue (WAT) acts as the main storage depot for energy, in the form of TAG rich lipid
droplets, which comprise approximately 90% of the total volume of the adipocyte (Xu et al., 2013).
Under times of low energy balance, such as physical exercise, WAT is charged with releasing
stored fatty acids (FA), via lipolysis of TAG, into circulation for use by the SKM and cardiac
muscle for fuel. Beyond providing a source of energy, adipocytes have been confirmed to secrete
their own cytokines, termed “adipokines.” The discovery of adipokines brings to light on a new
regulatory role for adipose tissue as a secretory tissue. Many adipokines have been shown to
activate the JAK/STAT signaling pathways throughout the body, such as leptin, TNF-α, and IL­
6, among others.
Recent discovery of active brown adipose tissue (BAT) in adult humans has launched several
studies to examine the molecular mediators of BAT expansion and activity (Virtanen et al., 2009).
BAT differs from WAT in that it is specialized in generating heat through non-shivering thermo-
genesis. Released FAs, via lipolytic processes, are required and metabolized in the mitochondria to
activate the BAT-specific protein UCP-1 (Cannon and Nedergaard, 2004). These two adipose
tissues (WAT and BAT) can be distinguished from one another other based on many features
(Frontini and Cinti, 2010). BAT is highly vascularized, has multi-locular lipid droplets, and
contains many mitochondria and myoglobin, whereas WAT lacks large amounts of vasculariza­
tion, is characterized by containing one unilocular lipid droplet, and few houses of mitochondria.
Interestingly, during chronic cold exposure, WAT has the ability to be “transdifferentiated” to
a BAT-like tissue, termed “browning” (Frontini and Cinti, 2010). While BAT is ample in new­
borns, the majority converts to WAT in adult humans. However, the mechanism inducing the
transformation between BAT and WAT is not well understood (Frontini and Cinti, 2010). All of
these aforementioned characteristics provide BAT with the ability to oxidize a large amount of FA
within the mitochondria. The discovery of active BAT in adult humans has led to a massive
amount of studies aimed at increasing either content or activity of the tissue for the potential to
treat obesity.

8.2.2 Adipose Tissue Contribution to Diabetes


Whole body insulin resistance is a hallmark precursor to the development of T2D. Of interest, the
adipose tissue is one of the first tissues to become resistant to insulin, thus reducing its ability to
take in glucose for fuel utilization (Bódis and Roden, 2018). In this regard, insulin resistance results
in an increase in lipolytic activity and a subsequent increase in circulating FAs. In turn the increase
in circulating FAs induce insulin resistance in other tissues, such as the liver and SKM. Further
consequence of the onset of insulin resistance is an increase in adipose tissue inflammatory
processes, that which induces insulin resistance in liver and SKM. An additional method by
which expansion of adipose tissue depots contributes to the development of T2D is through its
inability to take in additional post-prandial circulating FAs. Another theory for the contribution of
increased circulating FAs is that the lipid droplets within the adipocytes become “filled” with
TAGs. In this context, there is no additional capacity for TAG formation and storage within the
adipocyte. Many studies have been carried out analyzing the effects of various JAK/STAT actions
in adipose tissue (Figure 8.2).
Diabetes and Obesity 127

FIGURE 8.2 The intracellular role of JAK/STAT signaling in both white and BAT.

8.2.3 JAKs and White Adipocytes


There are many circulating activators of JAK enzymes in various cell types. In adipocytes the
major players, among others, appear to be IL-4, IL-6, GP-130, IFN-γ, Leptin, LIF, OSM, GH,
and PRL (Dodington et al., 2018; Zhao and Stephens, 2013). JAKs have been widely studied in
numerous cell types; however, there is little evidence suggesting that JAKs play a direct role in
mediating adipogenesis, independent from STAT interaction. Nevertheless, JAK1 and JAK2 have
been confirmed to be expressed and active in both adipose tissue and isolated adipocytes (Figure
8.2). JAK3 and Tyk2 have been reported to be expressed in adipose tissue, but not in adipocytes
themselves. Therefore, it is unknown whether or not JAK3 and/or Tyk2 play a significant role in
mediating metabolic processes specifically in WAT adipocytes. In WAT, JAK1 is one of the many
enzymes responsible for mediating insulin-stimulated glucose transport into the cells (McGilli­
cuddy et al., 2009). Additionally, JAK2 activity has been determined to be required for maintenance
of insulin sensitivity in WAT adipocytes (Shi et al., 2014). Alongside its insulin sensitizing action,
JAK2 is a mediator of lipolysis in adipose tissue, the systematic breakdown of stored triglycerides into
free fatty acid and glycerol. HFD-induced obesity is elevated in JAK2 knockout mice, potentially due
to the compromised ability to hydrolyze TAGs (Shi et al., 2016). On the other hand, adipocyte JAK2
expression increases as adipogenesis progresses and removal of JAK2 inhibits adipogenesis in cultured
adipocytes (Zhang et al., 2010). A unique feature of JAK2 action in white adipocytes is its ability to
bind, independently of cytokine activation and downstream phosphorylation of a STAT, to the fatty
acid binding protein, aP2 (Thompson et al., 2009). However, the implications of the interactions
between JAK2/aP2 have not been well established. Although JAK2 shows promise in its regulation
of pathways associated with fat storage and use, human studies have failed to link mutations in the
JAK2 gene to obesity (Ge et al., 2008). Similar to JAK2, JAK3 is required for maintenance of
“normal” body weight (Mishra et al., 2015). Further, when placed on high fat diet, mice lacking
JAK3 gain significantly more weight compared to their wild type counterparts.

8.2.4 STATs and White Adipocytes


Studies show that many of the STATs have the ability to reduce lipid deposition, through
induction of lipolysis in most cases, in fully differentiated mature adipocytes, while stimulating
128 JAK-STAT Signaling in Diseases

adipogenesis in pre-adipocytes. STAT1, 3, 5A and B expression levels increase in adipocytes with


differentiation, suggesting that they play an instrumental role in adipocyte function. Both
STAT5A and STAT5B have garnered a great deal of focus for their participation in promotion
and/or inhibition of adipogenesis. In adipocyte precursor cells, both STAT5 isoforms act to
promote adipogenesis while in mature adipocytes STAT5s play an anti-lipogenic role. STAT5
proteins become activated, via tyrosine phosphorylation, early on in the adipogenesis process in
cultured adipocytes (Stephens et al., 1996). Clear evidence for the role of STAT5 in promoting
adipogenesis has been demonstrated by the ability of both STAT5 isoforms to bind to pro­
adipogenic factors PPARγ and C/EBPβ and induce their transcriptional activity (Kawai et al.,
2007). Additionally, when STAT5A or STAT5B are overexpressed in preadipocytes, adipogenesis
is stimulated (Wakao et al., 2011). A unique model was utilized to overexpress STAT5A in
cultured 3T3 adipocytes which were subsequently injected into athymic mice. Fat pads developed
at the site of injection in the cells expressing the virally induced STAT5A (Stewart et al., 2011).
On the other hand, in mature adipocytes STAT5s have been shown to be an anti-lipogenic
intermediary of growth hormone action (Richard and Stephens, 2014). STAT5A has been shown
to promote adipogenesis in cells not programmed to become adipocytes (Floyd and Stephens,
2003). Of note, in a study of polymorphisms of the PPAR γ3 promoter, a C/G polymorphism was
detected within a STAT5B consensus site. The polymorphism site is of importance as it has been
associated with an increase in height and low-density lipoprotein cholesterol levels in humans
(Meirhaeghe et al., 2002).
The relevance of STAT6 in mediating metabolic processes in adipose tissue remains unclear,
although its expression is maintained throughout adipocyte differentiation, but its activity has been
reported to be reduced in mature adipocytes when compared to preadipocytes (Deng et al., 2000).
STAT1 has been shown to be the downstream STAT for the JAK1-mediated insulin signaling
pathway stimulating glucose uptake (McGillicuddy et al., 2009). STAT1 has also been suggested to
be essential for the lipolytic and antiadipogenic effects of JAK1. Further, STAT1 has the ability to
bind to the adipogenic transcriptional regulator, PPARγ, to mediate adipogenesis. However, whole
body STAT1 knockout mice do not display an obese phenotype (Meraz et al., 1996). Additional
studies have shown that STAT1 may promote adipogenesis. STAT3 has also been implicated in
being essential for the maintenance of body weight as loss of STAT3 in adipocytes results in an
obese phenotype with no disruption in energy expenditure (Cernkovich et al., 2008). On the other
hand, inhibition of STAT3 results in increased adipogenesis (Deng et al., 2006) and STAT3 has
been confirmed to induce pre-adipocytes to differentiate and proliferate into mature adipocytes
(Wang et al., 2012; Zhang et al., 2010). Overall, there appears to be a difference between STAT
function in cultured cells and in whole body in vivo mouse models.

8.2.5 JAKs and Brown Adipocytes


Even though it is not clear if Tyk2 is present in WAT cells, BAT cells do express Tyk2, and its
content is reduced when obesity is induced by high fat diet in mice. Further confirmation for the
role of Tyk2 in preventing obesity can be seen when mice that are lacking whole body Tyk2
display an obese phenotype. This obese phenotype is due to reductions in whole body energy
metabolism because the BAT is dysfunctional when Tyk2 is not expressed (Derecka et al., 2012).
JAK2 is another JAK that acts to induce BAT function and in turn increases whole body
metabolism to prevent obesity (Shi et al., 2016). Not only is JAK2 involved in metabolic
processes, there are higher levels of JAK2 expressed in BAT when compared to various WAT
depots. Further confirmation for the importance of JAK2 in mediating BAT function is that in
adipose-specific JAK2, knockout mice thermogenesis is impaired (Shi et al., 2016).

8.2.6 STATs and Brown Adipocytes


In order for JAKs to exert their effects on brown adipocytes, downstream activation of STATs is
required. STAT3 has been established as a necessary factor for brown adipocyte differentiation and it
Diabetes and Obesity 129

appears that its upstream JAK is Tyk2, required for its effects in brown adipocytes. Beyond a direct role
for STAT3 on brown adipocyte function, deletion of STAT3 in neuronal cells resulted in reductions in
BAT function and reduced thermogenesis (Gao et al., 2004). Beyond STAT3, there is little evidence that
alternative STATs play a role in intracellular brown adipocyte function and it is not known which JAK/
STAT combination is needed to increase brown adipocyte activity and content (Richard and Stephens,
2014; Shi et al., 2016; Ye, 2015). Nevertheless, STAT5 has been found to mediate thermogenesis through
indirect regulation (Lee et al., 2008). Upon STAT5 deletion in the CNS, the mice became severely obese
and were not able to maintain temperature homeostasis (Lee et al., 2008). Similarly, mice lacking STAT6
showed reduced core body temperature and a reduced thermogenic phenotype (Nguyen et al., 2011).
The authors concluded that intracellular STAT6 in macrophages, residing in BAT, may be responsible
for stimulating brown adipocyte function through norepinephrine signaling. Norepinephrine is a known
and potent activator of “browning” of white adipocytes and may serve as a link between STAT signaling
and induction of BAT (Horwitz et al., 1969).

8.3 Skeletal Muscle and Obesity


The SKM is the largest energy expending tissue in the body. Its major function is to coordinate
movement through muscle contraction. In order to carry out muscle contraction, an exorbitant
amount of energy is required through metabolizing glucose and FA, primarily. Therefore, targeting
the growth and development of lean mass is of interest in regards to weight loss. In obese states,
intramuscular TAG stores accumulate and mitochondrial dysfunction ensues. Further, in insulin-
resistant states, as seen with T2D, the muscle is unable to take in glucose and/or fat for fuel, leading
to further increases in circulating glucose levels. Many interventions have focused on increasing
physical activity to induce fat metabolism and prevent or treat insulin resistance (Delahanty, 2002;
Tuomilehto et al., 2001). However, behavior change interventions have not been successful for
maintaining weight loss in adults and children (Miller and Dunstan, 2004; Nooijen et al., 2017). As
a result of these failures, many researchers are attempting to elucidate signaling pathways responsible
for regulating fuel use. Recently, like adipose tissue, SKM has been implicated in being a secretory
organ with the ability to secrete numerous cytokines, termed “myokines,” in response to exercise
(Pedersen and Febbraio, 2012). Secretion of these cytokines is of interest as many exert their actions
via paracrine and/or autocrine manners to regulate metabolism. Myokines, such as interleukins are
important players as they activate JAK/STAT signaling (Ye, 2015). Because SKM is an energy
expending tissue, increases in SKM mass can lead to increases in fat utilization and reductions in
obesity. JAK/STAT pathways have been implicated in mediating intracellular events leading to
induction of SKM myogenesis (Jang and Baik, 2013). Disruptions in SKM JAK/STAT signaling
may lead to reductions in intracellular metabolism as well as whole body metabolism (Figure 8.3).

8.3.1 JAK Signaling in Skeletal Muscle


There is little information regarding the role of JAKs in moderating SKM metabolic processes,
especially in mouse models. However, in humans with T2D, SKM biopsies show increased
phosphorylated JAK2 (Mashili et al., 2013). Additional studies in obese humans revealed reduced
expression of Tyk2 in SKM biopsies (Derecka et al., 2012). Further, JAK3 has been confirmed to
be involved in inducing glucose uptake, via GLUT4 translocation to the plasma membrane, in
SKM myotubes (Krolopp et al., 2016). The lack of available information regarding JAK signaling
in SKM points to an area of much needed research.

8.3.2 STAT Signaling in Skeletal Muscle


STAT3 has been one of the most widely studied STAT in SKM, with contradictory findings in
regards to its role in mediating insulin signaling and glucose tolerance (Cui et al., 2004; Krolopp
130 JAK-STAT Signaling in Diseases

FIGURE 8.3 The intracellular role of JAK/STAT signaling in skeletal muscle.

et al., 2016; Wegrzyn et al., 2009). Some studies have proposed that STAT3 induces insulin-resistant
states via activation by IL-6, a confirmed activator of insulin resistance, and inhibition of STAT3 in
human insulin-resistant myotubes restored insulin sensitivity and/or prevented lipid-induced insulin
sensitivity (Kim et al., 2013; Mashili et al., 2013; Weigert et al., 2004). In line with these findings,
phosphorylation of STAT3 in SKM myotubes obtained from T2D humans is increased (Mashili
et al., 2013). On the other hand, STAT3 has been implicated in promoting glucose uptake in SKM
myotubes, while knock out of STAT3 in mice did not prevent high fat diet-induced insulin resistance
(Krolopp et al., 2016; White et al., 2015). Of note, SKM-specific knockout of STAT3 did not alter
energy expenditure; however, the mice displayed a glucose intolerance phenotype (White et al., 2015).
Interestingly, STAT5 appears to play a unique role in SKM, as it has been found to be involved in
maintaining lean mass via growth hormone-mediated activation (Klover et al., 2009; Klover and
Hennighausen, 2007). Loss of SKM STAT5 in mice has also been found to promote accumulation of
lipids in SKM, presumably via increased FA transport into the myotubes (Baik et al., 2017).

8.4 Pancreas and Diabetes


Pancreatic β-cells, housed within the islets of Langerhans, are highly specialized cells and respond to
increases in circulating blood glucose by releasing insulin into the blood stream. Proper β-cell
function is required to prevent the onset of both T1D and T2D. As aforementioned, increased body
fat results in dysfunctional β-cell response to alterations in glucose homeostasis. This provides
a direct link between obesity and diabetes. Further the development of T2D has been associated
with inflammatory-induced β-cell dysfunction (Donath et al., 2009). Since obesity is considered to
be a state of constant low-grade inflammation, this is an alternative role for the induction of diabetes
associated with body fatness. Importantly, insulin secretion is a tightly regulated process with many
intracellular signaling pathways playing a role, owing to the complexity of β-cell function.
The concentration of circulating glucose is highly structured and is not subject to large
fluctuations in healthy individuals. Pancreatic β-cells detect increases in blood glucose levels, as
Diabetes and Obesity 131

seen with postprandial hyperglycemia, and respond by releasing granules of insulin into the
circulation. Once in circulation, insulin binds cell-membrane insulin receptors in hepatocytes,
myocytes and adipocytes, where it drives increased glucose uptake in these cells to restore
euglycemia. De-regulation of this process results is a culprit in the development of T2D. In
order to study the role of JAK/STAT signaling in pancreatic function (Figure 8.4), many
genetically altered mouse models have been developed in specifically in β-cells. Findings from
studies utilizing these knockout models are of interest, due to the whole-body metabolic effects
seen from dysfunctional signaling in the β-cells. However, as will be discussed, findings from β-cell
specific knockout mice must be analyzed cautiously (Figure 8.4).

8.4.1 JAK Signaling in β-cell Function


Of the JAK family, JAK1, JAK3, and Tyk2 have been studied and implicated in mediating β-cell
function. β-cell–specific knock out of JAK2 revealed little effects on overall metabolism with no
alterations in overall body weight under high fat feeding conditions. It was hypothesized that
deletion of JAK2 would result in a reduction of β-cell mass; however, no alterations in β-cell mass
and/or expansion were observed (Choi et al., 2011). These aforementioned findings were unexpected
as JAK2 signaling has been implicated in mediating the effects of growth hormone in β-cells (Liu
et al., 2004). Although loss of JAK3 has been shown to induce insulin resistance, its role in β-cell
mass and/or function has not been examined. Further, Tyk2 is an essential JAK in the maintenance
of glucose tolerance, yet these observations have not been examined on the pancreatic level. Of
interest, β-cell mass was found to be increased in mice that contained liver-specific deletion of
JAK2, suggesting a “cross talk” between tissues (Shi et al., 2012). Although many of the JAKs have
been implicated in mediating process associated with β-cell function, such as glucose sensitivity and
insulin secretion, studies are limited in regards to direct examination on the β-cells.

8.4.2 STAT Signaling in β-cell Function


Unlike the JAKs, many STATs have been examined in β cell-specific mouse models. On the other
hand, little is known about the role of STATs to mediate insulin secretion in β-cells. Nevertheless,
β cell-specific knockout of STAT3 proved to disrupt many metabolic processes. Indeed, an increase

FIGURE 8.4 The intracellular role of JAK/STAT signaling in pancreas.


132 JAK-STAT Signaling in Diseases

in body weight, fat mass, and fasting blood glucose, with subsequent reductions in glucose tolerance
was observed in these aforementioned mouse models (Cui et al., 2004). Additional studies aimed at
examining the essential role of STAT3 in pancreatic function revealed loss of physiological first
phase insulin secretion from β-cells, along with aberrations in the morphology of the islets
(Gorogawa et al., 2004). Similar to STAT3, knockout of STAT5 in β-cells increased fat mass and
fasting blood glucose with reductions in glucose tolerance. Although there was an increase in fat
mass and circulating FAs, in the STAT5 β-cell specific knockout mice, there were no alterations
found on overall body weight (Mziaut et al., 2006).
Instead of knockdown of STAT5, alternative studies have been carried out in a model of
overexpression of dominant negative STAT5, and similar results were observed as in the β-cell­
specific model, with high fat feeding (Jackerott et al., 2006). Novel data from over-expression of
a dominant negative STAT5 isoform resulted in reductions in β-cell proliferation. Paradoxically,
islets isolated from the dominant negative mice displayed no defects in leptin-stimulated insulin
secretion. On the other hand, the investigators also examined the phenotype of mice over
expressing active STAT5 in β-cells. The genetically manipulated increase in STAT5, within the
β-cells, had no effects on body weight or fasting blood glucose levels, but did restore overall
glucose tolerance (Jackerott et al., 2006).
Of note, these previously mentioned β-cell-specific knockout studies were carried out by
knocking out the genes of interest using the rat-insulin promoter (RIP). Conclusions drawn
from studies using the RIP promoter must be drawn with caution, as this promoter has been
determined to be “leaky,” resulting in reductions in hypothalamic expression of genes of interest.
As will be discussed in a later section, the hypothalamus is a target for JAK/STAT signaling and
plays a role in maintain body weight and adiposity. To counteract the potential knockout of genes
in the hypothalamus, utilizing the RIP promoter, additional genetic studies have been performed
utilizing an alternative promoter, expressed in both endocrine and exocrine pancreas tissue
(PDX). In this context, both STAT3 and STAT5 were conditionally knockout in β-cells by
different researcher groups. Knockdown of STAT3, via the PDX promoter, yielded similar
results as the previous studies with reductions in glucose tolerance and glucose-stimulated
insulin secretion (Kostromina et al., 2010). However, the STAT3-PDX mice did not display
alterations in body weight, unlike the STAT3-RIP mice, suggesting that the hypothalamus might
play a role in the previously observed phenotype. Likewise, deletion of STAT5, on the PDX
promoter, did not replicate the STAT5-RIP model with having no measurable alterations in
body weight, glucose tolerance, or islet formation (Lee et al., 2007). Overall, the physiological
role of STAT3 and STAT5 in β-cells does not appear to be involved in mediating whole body
metabolism.
In vitro studies have shown that both STAT3 and STAT6 increased activity, as indicated by
increased phosphorylation, in human islets and in INS cells, as a result of activation of pro-
inflammatory cytokines. The results of these aforementioned studies suggest that both STAT3 and
STAT6 are involved in orchestrating the detrimental effects of inflammatory processes on β-cell
function. Independent of inflammation, the induction of STAT3 and 6 phosphorylation reduced cell
viability providing further confirmation that STATs play a negative role in maintaining β-cell function
(Russell et al., 2013).

8.5 Liver and Metabolic Processes


The liver plays a central role in mediating many metabolic processes as well as acting as
a detoxification site. The liver has the unique ability to regenerate following injury, owing to its
central importance for survival. Sustained obesity results in defects in liver function for a variety of
reasons. Defects in liver function disrupt many metabolic processes such as glucose homeostasis.
Non-alcoholic fatty liver disease (NAFLD) is a primary dysfunction occurring due to sustained
obesity. High fat diet and a sedentary lifestyle are major risk factors for the development of
Diabetes and Obesity 133

NAFLD. Hepatic steatosis is the severe deposition of fat within the liver and is strongly associated
with obesity and can stem from NAFLD. Preventing hepatic steatosis is of utmost importance as
the condition can lead to fibrosis and hepatocellular cancer. Further, excess fat deposition in
hepatocytes may result in local cytotoxicity and oxidative stress. In regards to metabolic processes,
physiological hepatocyte function is necessary for the control of lipid metabolism and storage and
exportation of lipoproteins. An additional important role of the liver is to maintain glucose
homeostasis, by storage and release of glucose into circulation (Figure 8.5).

8.5.1 JAK Signaling in Liver Metabolism


Of the numerous ligands associated with activated JAK/STATs, IL-4, IL-6, and GH appear to
be the most relevant in regard to liver metabolism. Knockout of JAK2 in the liver reduces body
weight and adiposity in mice while on high fat diet. However, these mice displayed hepatic
steatosis with hyperlipidemia, alongside increases in circulating FAs, and cholesterol levels.
These findings suggest that because the adipose tissue failed to expand, the liver then took on
the excess dietary fat, which corresponds to the increase in the fatty acid transporter, CD36, in
the livers of these mice (Sos et al., 2011). In additional studies examining the consequences of
JAK2 deletion in the liver, high fat diet-induced glucose intolerance was prevented (Shi et al.,
2012). Alongside the maintenance of glucose homeostasis in these mice, their energy expendi­
ture was higher, which explains the reduced body weight. Livers collected from mice lacking
JAK3 are larger and contain more fat deposition (Mishra et al., 2015). However, JAK3 was
deleted in the context of a whole body; therefore, it is not clear if the observed liver dysfunction
is locally mediated via JAK3.

8.5.2 STAT Signaling in Liver Metabolism


Of the STAT proteins, STAT3 is the most widely studied in regulating liver function. Interestingly,
STAT3 becomes phosphorylated with infusion of both glucose and insulin, independently, in
hepatocytes. Additionally, STAT3 has been identified as a protective factor in the development of

FIGURE 8.5 The intracellular role of JAK/STAT signaling in liver.


134 JAK-STAT Signaling in Diseases

fatty liver disease when associated with an upstream cytokine, IL-22 (Wang et al., 2011). Various
mouse models have been developed to conditionally delete STAT3 from hepatocytes. Insulin
resistance is a resulting factor stemming from liver-specific deletion of STAT3 (Inoue et al., 2004).
An additional phenotype observed in mice lacking STAT3 in liver, is an increase in gluconeogenic
genes. These findings were replicated in in vitro studies as well, owing to the importance of STAT3
in maintaining liver glucose metabolism (Inoue et al., 2004). Interestingly, when STAT3 was
overexpressed in the liver, via adenoviral injection, diabetic mice displayed reduced blood glucose,
plasma insulin concentrations, and hepatic gluconeogenic gene expression. An alternative conditional
knockout of STAT3 in liver yielded similar results as other STAT3 knockout mice, in that insulin
sensitivity was reduced (Moh et al., 2008).
STAT5 is also a major player in maintaining liver homeostasis. With conditional knockout of
STAT5 in the liver, mice develop hepatic steatosis, display lower body weight, increased circulating
insulin levels, and reduced insulin sensitivity (Cui et al., 2007). As with JAK2 liver-specific deletion,
CD36 expression increased as a result of loss of STAT5 in the liver, providing a mechanistic
explanation for the increased lipid accumulation in the liver (Hosui et al., 2016). Increased lipid
accumulation was observed in double knockout mice (STAT5 and CD36), confirming the role of
CD36 in mediating STAT5-induced hepatic steatosis (Hosui et al., 2016). STAT1 and STAT3 were
subsequently activated in the liver with removal of STAT5, suggesting a redundant role for STAT1,
3, and 5 (Cui et al., 2007).
Because the liver is a heterogeneous tissue, it is difficult to determine the cell types that are
responsible for various metabolic stimuli. Recent studies aimed at examining the novel role of
STAT6 in liver function were carried out in isolated primary hepatocytes (Ricardo-Gonzalez
et al., 2010). STAT6 phosphorylation was induced when its upstream cytokine activator, IL-4, was
applied to the cells, verifying the capability of STAT6 to be activated in hepatocytes. A further
consequence on liver function, with the removal of STAT6, was a shift towards FA use over
glucose use in the hepatocytes (Ricardo-Gonzalez et al., 2010).

8.6 Central Nervous System and Obesity


The Central Nervous system is the hub for control of food intake. In turn the hypothalamus is the
major region within the brain that is responsible for the satiety response and cessation of food
intake (Morton et al., 2006). A major mediator of food intake is leptin, which has been
characterized as an important adipokine responsible for stimulating satiety via activation of
expression of genes acting on the hypothalamus to reduce food intake (Myers et al., 2008). With
sustained high fat diet, the hypothalamus can become resistant to both insulin and leptin, thus
reducing the sensitivity of the insulin signaling pathway and reducing the satiety effects of leptin,
contributing to the development of obesity and ultimately T2D (De Souza et al., 2005; Morton
et al., 2006; Münzberg et al., 2004).
Since the discovery of leptin, many of its targets, downstream of its receptor, have been
identified. JAK/STAT signaling is one such pathway implicated in mediating the effects of leptin
on food intake (Figure 8.5) (Bjørbæk et al., 1997). Beyond the hypothalamus, the ventral
tegmental area has also been identified as an area within the CNS responsible for controlling
food intake (Morton et al., 2009). Undoubtedly, the CNS is a key player in controlling obesity
through the identification of signals signaling to increase or reduce food intake (Figure 8.6).

8.7 JAKs and Food Intake


As previously described, JAK/STATs are clearly involved in regulating intracellular metabolism in
various cell types to regulate obesity. An alternative role for JAK/STATs, in the regulation of
metabolic signals, is to modulate food intake via leptin action in the neurons within the
Diabetes and Obesity 135

FIGURE 8.6 The intracellular role of JAK/STAT signaling in the central nervous system.

hypothalamus (Farooqi and O’Rahilly, 2014). Upon leptin binding to its receptor, JAK2 phos­
phorylates and sets into motion activation of the insulin receptor substrate signaling pathway to
mediate metabolic processes within the neurons of the hypothalamus (Farooqi and O’Rahilly,
2014). Additionally, JAK2 plays a role in mediating leptin signals in the ventral tegmental area to
control food intake (Morton et al., 2009).

8.8 STATs and Food Intake


Importantly, with JAK2 activation STAT3 is recruited, phosphorylated, and dimerized for its
translocation to the nucleus, leading to increased expression of pro-opiomelanocortin (POMC)
(Bates et al., 2003). POMC is a player in signals that act on centers within the paraventricular
hypothalamus to regulate food intake. Mutations in genes associated with POMC signaling have
been verified to contribute to obesity. Beyond activation via leptin, STAT3 has also been shown to
be activated by two other cytokines involved in controlling food intake, TNFα, and ciliary
neurotrophic factor, via hypothalamic signals (Ladyman and Grattan, 2013). Further evidence
for neuronal STAT3 mediation of obesity is evident by STAT3 deletion in neuronal cells in mice,
resulting in an obese and diabetic phenotype (Gao et al., 2004). In addition to a severe increase in
fat mass, mice lacking STAT3 in neuronal cells also displayed disruptions in thermogenesis which
was attributed to reduced BAT (Gao et al., 2004). Further, when STAT3 is inactivated in the
hypothalamus, leptin loses its ability to maintain proper glucose homeostasis as mediated by liver
function (Buettner et al., 2006). CNS STAT5 has also been implicated in controlling food intake
as shown by increased food intake in deletion of STAT5 from CNS cells (Lee et al., 2008). When
STAT5 is deleted in the CNS, POMC expression is reduced, which may lead to the increase in
food intake.

8.9 Role of STATs in Mitochondrial Function


Although controversial, many of the STAT family of transcription factors (STAT1, STAT2,
STAT3, STAT5 and STAT6) have been shown to be associated with the mitochondria (Kramer
et al., 2015; Meier and Larner, 2014). Mitochondria have often been referred to as “power
houses” of the cell. Mitochondria produce ATP through oxidative phosphorylation to generate
energy for maintenance of cellular processes. In metabolically active tissues, such as SKM,
mitochondrial function plays a large role in mediating energy homeostasis. Alternatively, mitochon­
dria are responsible for “burning” fuel, FAs, and glucose, to reduce energy stores in the adipose tissue.
Therefore, a strategy to reduce stored fat depots is to induce increases in mitochondrial activity in all
tissues, especially SKM.
136 JAK-STAT Signaling in Diseases

Traditional, non-pharmacological methods to increase mitochondrial activity is through physical


exercise, subsequently increasing SKM mass (Yan et al., 2011). However, as noted previously,
maintenance of exercise routines is not an effective long-term strategy for weight management in
humans (Theofilou and Saborit, 2013). Therefore inducing molecular pathways mediating mito­
chondrial activity could be an alternative route to reduce fat mass. Because of the necessity for
energy production, mitochondrial processes are tightly regulated to maintain homeostasis. STAT3 is
the most widely studied STAT in regards to its effects on mitochondrial function. Indeed, STAT3
has been confirmed to be imported to the mitochondria and associate with complex I and II of the
electron transport chain, seemingly to increase activity of ATP generation (Tammineni et al., 2013;
Wegrzyn et al., 2009). Altogether, the study of STAT-mediated effects on mitochondrial activity and
function remains an open field.

8.10 Potential Pharmacological Therapies


There are many pharmacological inhibitors/activators of JAKs and STATs that have been used in
various experimental models. One such inhibitor, ruxolitinib, inhibits JAK1/2 and has been
studied in clinical trials for the treatment of certain cancers. Itacitinib, another JAK1 selective
inhibitor, has been used in clinical trials for cancer treatment as well (Beatty et al., 2018). But it is
not known if ruxolitinib and/or itacitinib has effects on mediating obesity and/or diabetes. It does
not seem likely, since knockout of JAK2 in adipose tissue results in an increase in fat expansion
and a reduction in energy expenditure. Extensive studies, aimed at discovering novel pharmaco­
logical JAK/STAT activators, have been carried out (Tai et al., 2012). Researchers screened over
1,400 naturally derived compounds and only four were identified (emodin, quercetin, apigenin
and luteolin) as having the ability to activate JAK/STAT signaling. However, it does not appear
that any of these compounds have been widely studied and certainly not in context related to
metabolic processes.
Recent studies have developed methodologies to screen small molecule inhibitors of STAT5B
(de Araujo et al., 2019). Identifying inhibitors of STAT5B is of importance as potential therapies
for obesity, as STAT5B is a stimulator of early adipogenesis. Conversely, STAT5B reduces
lipogenesis in mature adipocytes, which could be a potential downfall to utilizing inhibitors.
Another major hurdle to developing pharmacological mediators of JAK/STAT signaling is their
ubiquitous expression and often contradictory effects in various tissues, according to rodent
studies and in vitro studies.

8.11 Summary
It is inarguable that obesity is a growing and serious disease that which results in the progression
of many metabolic diseases, such as T2D. It is known that lifestyle alterations can counteract
obesity and its associated disease processes, but, as previously mentioned, humans have difficulty
maintaining lifestyle modifications for the long term. In this regard, many metabolic signaling
pathways have been studied for their potential to act as therapies for preventable diseases, such
as obesity. JAK/STAT signaling is a pathway that is under investigation for the treatment of
obesity and diabetes. Here it is shown that proper JAK/STAT signaling is relevant in various
tissues (adipose, SKM, liver, pancreas, hypothalamus) for mediating metabolic processes. It is
also clear that targeting JAK/STAT pathways for the treatment of obesity is a difficult endeavor,
owing to the complex and contradictory signaling in various tissues. Further, utilizing genetically
manipulated mouse models to examine signaling pathways does not always provide direct
translation to human models. In this regard, further studies examining the role of JAK/STATs
in a human context are warranted.
Diabetes and Obesity 137

REFERENCES
Baik, M., Lee, M. S., Kang, H. J., Park, S. J., Piao, M. Y., Nguyen, T. H., & Hennighausen, L. 2017.
Muscle-specific deletion of signal transducer and activator of transcription 5 augments lipid
accumulation in skeletal muscle and liver of mice in response to high-fat diet. European Journal of
Nutrition, 56(2), 569–579. doi:10.1007/s00394-015-1101-0.
Bates, S. H., Stearns, W. H., Dundon, T. A., Schubert, M., Tso, A. W., Wang, Y., … Myers, M. G. 2003.
STAT3 signalling is required for leptin regulation of energy balance but not reproduction. Nature, 421,
856–859.
Beatty, G. L., Shahda, S., Beck, T., Uppal, N., Cohen, S. J., Donehower, R., … Von Hoff, D. D. 2018.
A phase Ib/II study of the JAK1 inhibitor, itacitinib, plus nab-paclitaxel and gemcitabine in
advanced solid tumors. The Oncologist. doi:10.1634/theoncologist.2017-0665.
Biener, A., Cawley, J., & Meyerhoefer, C. 2017. The high and rising costs of obesity to the US health care
system. Journal of General Internal Medicine, 32(Suppl 1), 6–8. doi:10.1007/s11606-016-3968-8.
Bjørbæk, C., Uotani, S., Da Silva, B., & Flier, J. S. 1997. Divergent signaling capacities of the long and
short isoforms of the leptin receptor. Journal of Biological Chemistry, 272(51), 32686–32695.
doi:10.1074/jbc.272.51.32686.
Bódis, K., & Roden, M. 2018. Energy metabolism of white adipose tissue and insulin resistance in
humans. European Journal of Clinical Investigation, 0(0), e13017. doi:10.1111/eci.13017.
Buettner, C., Pocai, A., Muse, E. D., Etgen, A. M., Myers, M. G., & Rossetti, L. 2006. Critical role of
STAT3 in leptin’s metabolic actions. Cell Metab, 4(1), 49–60.
Cannon, B., & Nedergaard, J. A. N. 2004. Brown adipose tissue: Function and physiological significance.
Physiological Reviews, 84(1), 277–359. doi:10.1152/physrev.00015.2003.
Cernkovich, E. R., Deng, J., Bond, M. C., Combs, T. P., & Harp, J. B. 2008. Adipose-specific disruption of
signal transducer and activator of transcription 3 increases body weight and adiposity. Endocrinol­
ogy, 149(4), 1581–1590.
Choi, D., Cai, E. P., & Woo, M. 2011. The redundant role of JAK2 in regulating pancreatic β-cell mass.
Islets, 3, 389–392.
Cui, Y., Hosui, A., Sun, R., Shen, K., Gavrilova, O., Chen, W., … Hennighausen, L. 2007. Loss of signal
transducer and activator of transcription 5 leads to hepatosteatosis and impaired liver regeneration.
Hepatology, 46(2), 504–513. doi:10.1002/hep.21713.
Cui, Y., Huang, L., Elefteriou, F., Yang, G., Shelton, J. M., Giles, J. E., … Li, C. 2004. Essential role of
STAT3 in body weight and glucose homeostasis. Molecular and Cellular Biology, 24(1), 258–269.
doi:10.1128/MCB.24.1.258-269.2004.
Darnell, J. E. 1997. STATs and gene regulation. Science, 277(5332), 1630 LP–1635. Retrieved from http://
science.sciencemag.org/content/277/5332/1630.abstract.
de Araujo, E. D., Manaswiyoungkul, P., Erdogan, F., Qadree, A. K., Sina, D., Tin, G., … Gunning, P. T.
2019. A functional in vitro assay for screening inhibitors of STAT5B phosphorylation. Journal of
Pharmaceutical and Biomedical Analysis, 162, 60–65. doi:10.1016/j.jpba.2018.08.036.
De Souza, C. T., Araujo, E. P., Bordin, S., Ashimine, R., Zollner, R. L., Boschero, A. C., …
Velloso, L. A. 2005. Consumption of a fat-rich diet activates a proinflammatory response and
induces insulin resistance in the hypothalamus. Endocrinology, 146(10), 4192–4199. doi:10.1210/
en.2004-1520.
Delahanty, L. M. 2002. Evidence-based trends for achieving weight loss and increased physical activity:
Applications for diabetes prevention and treatment. Diabetes Spectrum, 15(3), 183 LP–189.
Retrieved from http://spectrum.diabetesjournals.org/content/15/3/183.abstract.
Deng, J., Hua, K., Caveney, E. J., Takahashi, N., & Harp, J. B. 2006. Protein inhibitor of activated STAT3
inhibits adipogenic gene expression. Biochemical and Biophysical Research Communications, 339(3),
923–931. doi:10.1016/j.bbrc.2005.10.217.
Deng, J., Hua, K., Lesser, S. S., Greiner, A. H., Walter, A. W., Marrero, M. B., & Harp, J. B. 2000.
Interleukin-4 mediates STAT6 activation in 3T3-L1 preadipocytes but not adipocytes. Biochemical
and Biophysical Research Communications, 267(2), 516–520. doi:10.1006/bbrc.1999.1993.
Derecka, M., Gornicka, A., Koralov, S. B., Szczepanek, K., Morgan, M., Raje, V., … Larner, A. C. 2012.
Tyk2 and Stat3 regulate brown adipose tissue differentiation and obesity. Cell Metabolism, 16(6),
814–824. doi:10.1016/j.cmet.2012.11.005.
138 JAK-STAT Signaling in Diseases

Dobrian, A. D., Galkina, E.V., Ma, Q., Hatcher, M., Aye, S. M., Butcher, M. J., … Nadler, J. L. 2013.
STAT4 deficiency reduces obesity-induced insulin resistance and adipose tissue inflammation.
Diabetes, 62(12), 4109–4121.
Dodington, D. W., Desai, H. R., & Woo, M. 2018. JAK/STAT – Emerging players in metabolism. Trends
in Endocrinology & Metabolism, 29(1), 55–65. doi:10.1016/j.tem.2017.11.001.
Donath, M. Y., Böni-Schnetzler, M., Ellingsgaard, H., & Ehses, J. A. 2009. Islet inflammation
impairs the pancreatic β-Cell in Type 2 diabetes. Physiology, 24(6), 325–331. doi:10.1152/
physiol.00032.2009.
Farooqi, I. S., & O’Rahilly, S. 2014. 20 YEARS OF LEPTIN: Human disorders of leptin action. Journal
of Endocrinology, 223(1), T63–T70. doi:10.1530/JOE-14-0480.
Floyd, Z. E., & Stephens, J. M. 2003. STAT5A promotes adipogenesis in nonprecursor cells and
associates with the glucocorticoid receptor during adipocyte differentiation. Diabetes, 52(2), 308
LP–314. Retrieved from http://diabetes.diabetesjournals.org/content/52/2/308.abstract.
Frontini, A., & Cinti, S. 2010. Distribution and development of brown adipocytes in the murine and
human adipose organ. Cell Metabolism, 11(4), 253–256. doi:10.1016/j.cmet.2010.03.004.
Gao, Q., Wolfgang, M. J., Neschen, S., Morino, K., Horvath, T. L., Shulman, G. I., & Fu, X.-Y. 2004.
Disruption of neural signal transducer and activator of transcription 3 causes obesity, diabetes,
infertility, and thermal dysregulation. Proceedings of the National Academy of Sciences of the United
States of America, 101(13), 4661 LP–4666. Retrieved from www.pnas.org/content/101/13/4661.abstract.
Ge, D., Gooljar, S. B., Kyriakou, T., Collins, L. J., Swaminathan, R., Snieder, H., … O’Dell, S. D. 2008.
Association of common JAK2 variants with body fat, insulin sensitivity and lipid profile. Obesity
(Silver Spring, Md.), 16(2), 492–496. doi:10.1038/oby.2007.79.
Gorogawa, S., Fujitani, Y., Kaneto, H., Hazama, Y., Watada, H., Miyamoto, Y., … Hori, M. 2004.
Insulin secretory defects and impaired islet architecture in pancreatic β-cell-specific STAT3 knock­
out mice. Biochemical and Biophysical Research Communications, 319(4), 1159–1170. doi:10.1016/j.
bbrc.2004.05.095.
Horwitz, B. A., Horowitz, J. M., & Smith, R. E. 1969. Norepinephrine-induced depolarization of brown
fat cells. Proceedings of the National Academy of Sciences of the United States of America, 64(1),
113–120. Retrieved from www.ncbi.nlm.nih.gov/pmc/articles/PMC286134/.
Hosui, A., Tatsumi, T., Hikita, H., Saito, Y., Hiramatsu, N., Tsujii, M., … Takehara, T. 2016. Signal
transducer and activator of transcription 5 plays a crucial role in hepatic lipid metabolism through
regulation of CD36 expression. Hepatology Research, 47(8), 813–825. doi:10.1111/hepr.12816.
Ihle, J. N., Nosaka, T., Thierfelder, W., Quelle, F. W., & Shimoda, K. 1995. Jaks and Stats in cytokine
signaling. Stem Cells, 15(S2), 105–112. doi:10.1002/stem.5530150814.
Inoue, H., Ogawa, W., Ozaki, M., Haga, S., Matsumoto, M., Furukawa, K., … Kasuga, M. 2004.Role of
STAT-3 in regulation of hepatic gluconeogenic genes and carbohydrate metabolism in vivo. Nature
Medicine, 10, 168. doi:10.1038/nm980.
Jackerott, M., Møldrup, A., Thams, P., Galsgaard, E. D., Knudsen, J., Lee, Y. C., & Nielsen, J. H. 2006.
STAT5 activity in pancreatic β-cells influences the severity of diabetes in animal models of type 1
and 2 diabetes. Diabetes, 55(10), 2705 LP–2712. Retrieved from http://diabetes.diabetesjournals.org/
content/55/10/2705.abstract.
Jang, Y.-N., & Baik, E. J. 2013. JAK-STAT pathway and myogenic differentiation. JAK-STAT, 2(2),
e23282. doi:10.4161/jkst.23282.
Kawai, M., Namba, N., Mushiake, S., Etani, Y., Nishimura, R., Makishima, M., & Ozono, K. 2007.
Growth hormone stimulates adipogenesis of 3T3-L1 cells through activation of the Stat5A/
5B-PPARgamma pathway. Journal of Molecular Endocrinology, 38, 19–34.
Kim, T. H., Choi, S. E., Ha, E. S., Jung, J. B., Han, S. J., Kim, H. J., … Lee, K. W. 2013. IL-6 induction of
TLR-4 gene expression via STAT3 has an effect on insulin resistance in human skeletal muscle. Acta
Diabetelogica, 50, 189–200.
Klover, P., Chen, W., Zhu, B.-M., & Hennighausen, L. 2009. Skeletal muscle growth and fiber composi­
tion in mice are regulated through the transcription factors STAT5a/b: Linking growth hormone to
the androgen receptor. The FASEB Journal, 23(9), 3140–3148. doi:10.1096/fj.08-128215.
Klover, P., & Hennighausen, L. 2007. Postnatal body growth is dependent on the transcription factors
signal transducers and activators of transcription 5a/b in muscle: A role for autocrine/paracrine
insulin-like growth factor I. Endocrinology, 148(4), 1489–1497.
Diabetes and Obesity 139

Kostromina, E., Gustavsson, N., Wang, X., Lim, C.-Y., Radda, G. K., Li, C., & Han, W. 2010. Glucose
intolerance and impaired insulin secretion in pancreas-specific signal transducer and activator of
transcription-3 knockout mice are associated with microvascular alterations in the pancreas.
Endocrinology, 151(5), 2050–2059. doi:10.1210/en.2009-1199.
Kramer, A. H., Kadye, R., Houseman, P. S., & Prinsloo, E. 2015. Mitochondrial STAT3 and reactive oxygen
species: A fulcrum of adipogenesis? JAK-STAT, 4(2), 1–10. doi:10.1080/21623996.2015.1084084.
Krolopp, J. E., Thornton, S. M., & Abbott, M. J. 2016. IL-15 activates the Jak3/STAT3 signaling pathway
to mediate glucose uptake in skeletal muscle cells. Frontiers in Physiology, 7, 626. doi:10.3389/
fphys.2016.00626.
Ladyman, S. R., & Grattan, D. R. 2013. JAK-STAT and feeding. JAK-STAT, 2(2), e23675. doi:10.4161/
jkst.23675.
Lee, J.-Y., Gavrilova, O., Davani, B., Na, R., Robinson, G. W., & Hennighausen, L. 2007. The transcription
factors Stat5a/b are not required for islet development but modulate pancreatic β-cell physiology
upon aging. Biochimica et Biophysica Acta (BBA) – Molecular Cell Research, 1773(9), 1455–1461.
doi:10.1016/j.bbamcr.2007.05.010.
Lee, J.-Y., Muenzberg, H., Gavrilova, O., Reed, J. A., Berryman, D., Villanueva, E. C., …
Hennighausen, L. 2008. Loss of cytokine-STAT5 signaling in the cns and pituitary gland
alters energy balance and leads to obesity. PLoS ONE, 3(2), e1639. doi:10.1371/journal.
pone.0001639.
Liu, J.-L., Coschigano, K. T., Robertson, K., Lipsett, M., Guo, Y., Kopchick, J. J., … Liu, Y. L. 2004.
Disruption of growth hormone receptor gene causes diminished pancreatic islet size and increased
insulin sensitivity in mice. American Journal of Physiology – Endocrinology and Metabolism, 287(3),
E405–E413. doi:10.1152/ajpendo.00423.2003.
Mashili, F., Chibalin, A. V., Krook, A., & Zierath, J. R. 2013. Constitutive STAT3 phosphorylation
contributes to skeletal muscle insulin resistance in type 2 diabetes. Diabetes, 62(2), 457 LP–465.
Retrieved from http://diabetes.diabetesjournals.org/content/62/2/457.abstract.
McGillicuddy, F. C., Chiquoine, E. H., Hinkle, C. C., Kim, R. J., Shah, R., Roche, H. M., … Reilly, M. P.
2009. Interferon γ attenuates insulin signaling, lipid storage, and differentiation in human adipocytes
via activation of the JAK/STAT pathway. Journal of Biological Chemistry, 284(46), 31936–31944.
doi:10.1074/jbc.M109.061655.
Meier, J. A., & Larner, A. C. 2014. Toward a new STATe: The role of STATs in mitochondrial function.
Seminars in Immunology, 26(1), 20–28. doi:10.1016/j.smim.2013.12.005.
Meirhaeghe, A., Fajas, L., Gouilleux, F., Cottel, D., Helbecque, N., Auwerx, J., & Amouyel, P. 2002.
A functional polymorphism in a STAT5B site of the human PPARγ3 gene promoter affects height
and lipid metabolism in a french population. Arteriosclerosis, Thrombosis, and Vascular Biology, 23,
289–294.
Meraz, M. A., White, J. M., Sheehan, K. C. F., Bach, E. A., Rodig, S. J., Dighe, A. S., … Schreiber, R. D.
1996. Targeted disruption of the Stat1 gene in mice reveals unexpected physiologic specificity in the
JAK–STAT signaling pathway. Cell, 84(3), 431–442. doi:10.1016/S0092-8674(00)81288-X.
Miller, Y. D., & Dunstan, D. W. 2004. The effectiveness of physical activity interventions for the treatment
of overweight and obesity and type 2 diabetes. Journal of Science and Medicine in Sport, 7(1), 52–59.
doi:10.1016/S1440-2440(04)80278-0.
Mishra, J., Verma, R. K., Alpini, G., Meng, F., & Kumar, N. 2015. Role of Janus kinase 3 in
predisposition to obesity-associated metabolic syndrome. The Journal of Biological Chemistry, 290
(49), 29301–29312. doi:10.1074/jbc.M115.670331.
Miyazaki, T., Kawahara, A., Fujii, H., Nakagawa, Y., Minami, Y., Liu, Z. J. et al. 1994. Functional
activation of Jak1 and Jak3 by selective association with IL-2 receptor subunits. Science, 266(5187),
1045 LP–1047.
Moh, A., Zhang, W., Yu, S., Wang, J., Xu, X., Li, J., & Fu, X.-Y. 2008. STAT3 sensitizes insulin signaling
by negatively regulating glycogen synthase kinase-3β. Diabetes, 57(5), 1227 LP–1221235. Retrieved
from http://diabetes.diabetesjournals.org/content/57/5/1227.abstract.
Morton, G. J., Blevins, J. E., Kim, F., Matsen, M., & Figlewicz, D. P. 2009. The action of leptin in
the ventral tegmental area to decrease food intake is dependent on Jak-2 signaling. American
Journal of Physiology – Endocrinology and Metabolism, 297(1), E202–E210. doi:10.1152/
ajpendo.90865.2008.
140 JAK-STAT Signaling in Diseases

Morton, G. J., Cummings, D. E., Baskin, D. G., Barsh, G. S., & Schwartz, M. W. 2006. Central nervous
system control of food intake and body weight. Nature, 443, 289. doi:10.1038/nature05026.
Münzberg, H., Flier, J. S., & Bjørbæk, C. 2004. Region-specific leptin resistance within the hypothalamus
of diet-induced obese mice. Endocrinology, 145(11), 4880–4889. doi:10.1210/en.2004-0726.
Myers, M. G., Cowley, M. A., & Munzberg, H. 2008. Mechanisms of leptin action and leptin resistance.
Annual Review of Physiology, 70, 37–56.
Mziaut, H., Trajkovski, M., Kersting, S., Ehninger, A., Altkrüger, A., Lemaitre, R. P., … Solimena, M.
2006.Synergy of glucose and growth hormone signalling in islet cells through ICA512 and STAT5.
Nature Cell Biology, 8, 435. doi:10.1038/ncb1395.
Nguyen, K. D., Qiu, Y., Cui, X., Goh, Y. P. S., Mwangi, J., David, T., … Chawla, A. 2011. Alternatively
activated macrophages produce catecholamines to sustain adaptive thermogenesis. Nature, 480
(7375), 104–108. doi:10.1038/nature10653.
Nooijen, C. F. J., Galanti, M. R., Engström, K., Möller, J., & Forsell, Y. 2017. Effectiveness of interventions
on physical activity in overweight or obese children: A systematic review and meta-analysis including
studies with objectively measured outcomes. Obesity Reviews, 18(2), 195–213. doi:10.1111/obr.12487.
Pedersen, B. K., & Febbraio, M. A. 2012. Muscles, exercise and obesity: Skeletal muscle as a secretory
organ. Nature Reviews Endocrinology, 8(8), 457–465. doi:10.1038/nrendo.2012.49.
Ricardo-Gonzalez, R. R., Eagle, A. R., Odegaard, J. I., Jouihan, H., Morel, C. R., Heredia, J. E., …
Chawla, A. 2010. IL-4/STAT6 immune axis regulates peripheral nutrient metabolism and insulin
sensitivity. Proceedings of the National Academy of Sciences, 107, 22617–22622.
Richard, A. J., & Stephens, J. M. 2014. The role of JAK-STAT signaling in adipose tissue function.
Biochimica et Biophysica Acta, 1842(3), 431–439. doi:10.1016/j.bbadis.2013.05.030.
Russell, M. A., Cooper, A. C., Dhayal, S., & Morgan, N. G. 2013. Differential effects of interleukin-13
and interleukin-6 on Jak/STAT signaling and cell viability in pancreatic β-cells. Islets, 5(2), 95–105.
doi:10.4161/isl.24249.
Schindler, C. W. 2002. Series Introduction: JAK-STAT signaling in human disease. The Journal of Clinical
Investigation, 109(9), 1133–1137. doi:10.1172/JCI15644.
Shi, S., Luk, C., Brunt, J., Sivasubramaniyam, T., Lu, S.-Y., Schroer, S., & Woo, M. 2014. Adipocyte-specific
deficiency of Janus kinase (JAK) 2 in mice impairs lipolysis and increases body weight, and leads to
insulin resistance with ageing. Diabetologia, 57(5), 1016–1026. doi:10.1007/s00125-014-3185-0.
Shi, S. Y., Martin, R. G., Duncan, R. E., Choi, D., Lu, S.-Y., Schroer, S. A., … Woo, M. 2012.
Hepatocyte-specific deletion of Janus kinase 2 (JAK2) protects against diet-induced steatohepatitis
and glucose intolerance. Journal of Biological Chemistry, 287(13), 10277–10288. doi:10.1074/jbc.
M111.317453.
Shi, S. Y., Zhang, W., Luk, C. T., Sivasubramaniyam, T., Brunt, J. J., Schroer, S. A., … Woo, M. 2016. JAK2
promotes brown adipose tissue function and is required for diet- and cold-induced thermogenesis in
mice. Diabetologia, 59(1), 187–196. doi:10.1007/s00125-015-3786-2.
Sos, B. C., Harris, C., Nordstrom, S. M., Tran, J. L., Balázs, M., Caplazi, P., … Weiss, E. J. 2011.
Abrogation of growth hormone secretion rescues fatty liver in mice with hepatocyte-specific
deletion of JAK2. The Journal of Clinical Investigation, 121(4), 1412–1423. doi:10.1172/JCI42894.
Stephens, J. M., Morrison, R. F., & Pilch, P. F. 1996. The expression and regulation of STATs during
3T3-L1 adipocyte differentiation. Journal of Biological Chemistry, 271(18), 10441–10444.
doi:10.1074/jbc.271.18.10441.
Stewart, W. C., Pearcy, L., Floyd, Z. E., & Stephens, J. M. 2011. STAT5A expression in Swiss 3T3 cells
promotes adipogenesis in vivo in an athymic mice model system. Obesity (Silver Spring, Md.), 19
(9), 1731–1734. doi:10.1038/oby.2011.66.
Tai, Z. F., Zhang, G. L., & Wang, F. 2012. Identification of small molecule activators of the Janus kinase/
signal transducer and activator of transcription pathway using a cell-based screen. Biological and
Pharmaceutical Bulletin, 35(1), 65–71. doi:10.1248/bpb.35.65.
Takeda, K., & Akira, S. 2000. STAT family of transcription factors in cytokine-mediated biological
responses. Cytokine & Growth Factor Reviews, 11(3), 199–207. doi:10.1016/S1359-6101(00)00005-8.
Tammineni, P., Anugula, C., Mohammed, F., Anjaneyulu, M., Larner, A. C., & Sepuri, N. B. V. 2013. The
import of the transcription factor STAT3 into mitochondria depends on GRIM-19, a component of
the electron transport chain. Journal of Biological Chemistry, 288(7), 4723–4732. doi:10.1074/jbc.
M112.378984.
Diabetes and Obesity 141

Theofilou, P., & Saborit, A. R. 2013. Adherence and physical activity. Health Psychology Research, 1(1),
e6–e6. doi:10.4081/hpr.2013.e6.
Thompson, B. R., Mazurkiewicz-Muñoz, A. M., Suttles, J., Carter-Su, C., & Bernlohr, D. A. 2009.
Interaction of adipocyte fatty acid-binding protein (AFABP) and JAK2: AFABP/aP2 as a regulator of
JAK2 signaling. Journal of Biological Chemistry, 284(20), 13473–13480. doi:10.1074/jbc.M900075200.
Tuomilehto, J., Lindström, J., Eriksson, J. G., Valle, T. T., Hämäläinen, H., Ilanne-Parikka, P., …
Uusitupa, M. 2001. Prevention of Type 2 diabetes mellitus by changes in lifestyle among subjects
with impaired glucose tolerance. New England Journal of Medicine, 344(18), 1343–1350.
doi:10.1056/NEJM200105033441801.
Virtanen, K. A., Lidell, M. E., Orava, J., Heglind, M., Westergren, R., Niemi, T., … Nuutila, P. 2009.
Functional brown adipose tissue in healthy adults. New England Journal of Medicine, 360(15),
1518–1525. doi:10.1056/NEJMoa0808949.
Wakao, H., Wakao, R., Oda, A., & Fujita, H. 2011. Constitutively active Stat5A and Stat5B promote
adipogenesis. Environmental Health and Preventive Medicine, 16(4), 247–252. doi:10.1007/s12199­
010-0193-7.
Wang, D., Zhou, Y., Lei, W., Zhang, K., Shi, J., Hu, Y., … Song, J. 2012. Signal transducer and activator
of transcription 3 (STAT3) regulates adipocyte differentiation via peroxisome-proliferator-activated
receptor γ (PPARγ). Biology of the Cell, 102(1), 1–12. doi:10.1042/BC20090070.
Wang, H., Lafdil, F., Kong, X., & Gao, B. 2011. Signal transducer and activator of transcription 3 in liver
diseases: A novel therapeutic target. International Journal of Biological Sciences, 7(5), 536–550.
Wegrzyn, J., Potla, R., Chwae, Y.-J., Sepuri, N. B. V., Zhang, Q., Koeck, T., … Larner, A. C. 2009.
Function of mitochondrial Stat3 in cellular respiration. Science, 323(5915), 793–797. doi:10.1126/
science.1164551.
Weigert, C., Brodbeck, K., Staiger, H., Kausch, C., Machicao, F., Häring, H. U., & Schleicher, E. D. 2004.
Palmitate, but not unsaturated fatty acids, induces the expression of interleukin-6 in human
myotubes through proteasome-dependent activation of nuclear factor-κB. Journal of Biological
Chemistry, 279(23), 23942–23952. doi:10.1074/jbc.M312692200.
White, A. T., LaBarge, S. A., McCurdy, C. E., & Schenk, S. 2015. Knockout of STAT3 in skeletal muscle
does not prevent high-fat diet-induced insulin resistance. Molecular Metabolism, 4(8), 569–575.
doi:10.1016/j.molmet.2015.05.001.
Xu, D., Yin, C., Wang, S., & Xiao, Y. 2013. JAK-STAT in lipid metabolism of adipocytes. JAK-STAT, 2,
e27203.
Yan, Z., Okutsu, M., Akhtar, Y. N., & Lira, V. A. 2011. Regulation of exercise-induced fiber type
transformation, mitochondrial biogenesis, and angiogenesis in skeletal muscle. Journal of Applied
Physiology (Bethesda, Md. : 1985), 110(1), 264–274. doi:10.1152/japplphysiol.00993.2010.
Ye, J. 2015. Beneficial metabolic activities of inflammatory cytokine interleukin 15 in obesity and type 2
diabetes. Frontiers of Medicine, 9(2), 139–145. doi:10.1007/s11684-015-0377-z.
Zhang, K., Guo, W., Yang, Y., & Wu, J. 2010. JAK2/STAT3 pathway is involved in the early stage of
adipogenesis through regulating C/EBPβ transcription. Journal of Cellular Biochemistry, 112(2),
488–497. doi:10.1002/jcb.22936.
Zhao, P., & Stephens, J. M. 2013. Identification of STAT target genes in adipocytes. JAK-STAT, 2(2),
e2309.
9
JAK-STAT Signaling in Liver Fibrosis

Marwa K. Ibrahim and Noha G. Bader El Din


Department of Microbial Biotechnology, Genetic Engineering and Biotechnology
Research Division
National Research Centre
Dokki, Egypt

9.1 Liver Fibrosis


Liver fibrosis is a multi-factorial pathological status resulting from excessive wound-healing process. The
etiologies include excessive alcoholic intake, chronic infections with hepatitis viruses such as hepatitis-C
virus (HCV) and-B virus (HBV), bile duct damage, and non-alcoholic fatty liver disease (NAFLD) or
non-alcoholic steatohepatitis (NASH) (Younossi et al. 2018). The causative factors lead ultimately to
liver fibrosis through induction of hepatocyte damage, uncontrolled proliferation of liver myofibroblasts,
hyper-activation of inflammatory cells in the liver milieu, and over-secretion of extracellular matrix
(ECM) components; which ends in the formation of fibrous scar. The presence of numerous fibrous scars
within the liver alters the overall architecture, disrupts liver function, causes hepatocyte loss, and
ultimately leads to liver failure (Kisseleva and Brenner 2008). Liver fibrosis can be reversed at its early
stages as long as it did not progress to advanced stages of cirrhosis (Brenner 2013; Seki and Brenner
2015). This fact sheds light on the worth of studying the molecular bases of the signaling pathways that
perform fibrotic activity. The latter will help discover new drugs with anti-fibrotic potentials.
Inflammation is a mutual driver of liver diseases, and the major trigger of liver tissue damage. The
inflammatory pathways play a significant role in the fibrogenesis process in liver. They lead to the
activation of hepatic stellate cells (HSCs), which then trans-differentiate into myofibroblasts that
produce, in tremendous amounts, ECM components (Akçalı and Aydın 2018). Of the various
inflammatory signaling pathways, the Janus kinase-signal transducer and activator of transcription
(JAK-STAT) pathway, in one hand, orchestrates the innate immune response in this organ to
different pathogens, and, on the other hand, plays a multitude of critical roles in the pathogenesis of
hepatic fibrosis. Signaling through JAK-STAT pathways results in a robust inflammatory response;
this notion makes this pathway as a great target for fibrosis hindrance.

9.2 JAK-STAT Pathway


The discovery of JAK-STAT pathway in mammals dates back to 1990s. The activation of JAK­
STAT pathway in the liver is mainly triggered by growth hormone, wide range of cytokines (Gao
2005), growth factors (Ruff-Jamison et al. 1993), and viral proteins (Machida et al. 2010).
Following hepatic injury, different cell subsets (Kupffer cells, HSCs, hepatocytes, natural killer
(NK) cells, dendritic cells, and lymphocytes) localized in the liver produce array of cytokines
either with inflammatory or hepatoprotective potentials. Of these cytokines, interferon-γ (IFN-γ),
IFN-α/β, interleukin-6 (IL-6), and IL-22 can activate JAK-STAT signaling pathway (Gao 2005).
In general, signaling cascade through JAK-STAT pathway is induced by the binding of the

143
144 JAK-STAT Signaling in Diseases

cytokines to their cognate receptors, which is followed by subsequent steps; receptor dimerization,
dimerization of receptor-associated JAK, a series of phosphorylation of JAKs and receptor, and
phosphorylation and activation of STAT proteins by the phosphorylated JAK–receptor complex.
These activated STATs dimerize (form homo-or heterodimers) and translocate into the nucleus, where
they induce the expression of genes involved in various cellular functions (Darnell et al. 1994; Schindler
1999). The family of JAKs encompasses of four non-receptor protein tyrosine kinases: JAK1, JAK2,
JAK3, and Tyk2 (Ghoreschi et al. 2009), while STATs include a set of seven proteins (STAT—1, 2, 3, 4,
5a, 5b, and 6). According to the structure of receptor intracellular domains, each cytokine receptor
activates specific group of individual JAKs and STATs (Gao et al. 2012). The activation of STATs by
certain cytokines plays a diverse array of functions in liver pathophysiology (Mair et al. 2011; Gao et al.
2012). In this chapter, the updated information on the implication of the effector arms of JAK-STAT
signaling pathway (STATs) in liver fibrogenesis is discussed at length in the below sections.

9.2.1 STAT1
STAT1 plays an important role in immune response and antiviral defense (Najjar and Fagard 2010;
Gao et al. 2012). STAT1 signaling in the liver is usually activated by IFN-α/β and IFN- γ. Also, STAT1
is activated in the liver by IL-27 cytokine, which is secreted by activated dendritic cells and macrophages
(Schroder et al. 2004). The activated STAT1 leads to the overexpression of many antiviral proteins. It
was reported that STAT1-deficient mice had increased sensitivity to pathogens infection because they
didn’t respond to interferon (Meraz et al. 1996). During the interferon treatment, the expression and
phosphorylation of STAT1 protein are increased in NK cells and induce killing of activated HSCs.
Also, the STAT1 elevation in NK cells contributes to NK cell cytotoxicity and antiviral activity (Edlich
et al. 2012). Moreover, IFN-λ (IL-28A and IL-28B, IL-29) causes prolonged STAT1 activation and
expression of many interferon stimulated genes (ISG), which provides protective effects against viral
infection and liver damage (Donnelly et al. 2011; Zhang et al. 2017, 2011).
In HSCs, STAT1 has anti-fibrotic effect by inhibiting the proliferation of HSCs, stimulating
apoptosis and cell cycle arrest (Jeong et al. 2006). The STAT1 upon phosphorylation (induced by
IFN-α) interacts with p300 and inhibits collagen gene transcription (Inagaki et al. 2003). Tanabe
et al. reported that the activation of STAT1 by IFN-β treatment reduces hepatic fibrosis in
concanavalin A (Con A)-induced hepatitis mice (Tanabe et al. 2007). IFN-γ has different anti­
fibrotic effects via stimulation of STAT1 phosphorylation, upregulation of mothers against
decapentaplegic homolog 7 (SMAD7) expression, and inhibition of transforming growth factor-
beta (TGF-β) (Weng et al. 2009; Tang et al. 2017). Additionally, IL-27/STAT1 pathway plays anti­
fibrotic role in HSCs cell line and inhibits liver fibrosis (Schroder et al. 2004). Jeong et al. reported
that STAT1-deficient mice, which are treated with carbon tetrachloride (CCl4), are more suscep­
tible to develop liver fibrosis and this finding demonstrated the protective effect of STAT1 against
liver fibrosis (Jeong et al. 2006). Recently, it has been shown that STAT1 activation stimulates
antitumor immunity and immune surveillance and STAT1 gene polymorphisms increase the risk
of developing hepatocellular carcinoma (HCC) (Zhu et al. 2007; Meissl et al. 2017).
On the other hand, several studies proved that STAT1 in hepatocytes induces the transcription of
several genes, which promote liver inflammation, injury, and fibrosis (Hong et al. 2002; Jaruga et al.
2003; Siebler et al. 2003; Wen et al. 2017). It was proved that the overexpression of STAT1 proteins
by IFN-γ develops chronic hepatitis in transgenic mice (Toyonaga et al. 1994). Also, the IFN-γ
induced cell death in hepatocellular carcinoma cell line Hep3B by overexpressing STAT1 protein.
Recent study demonstrated that targeting of IFN-γ/STAT1 signaling by T helper 1 (Th1) cells
accelerated the HSCs proliferation and secretion (Wen et al. 2017). Moreover, it was documented
that The STAT1 activation in hepatocytes is considered a pro-apoptotic signal, which increases cell
death and promotes liver damage (Hong et al. 2002; Gao et al. 2012). The disruption of STAT1
genes, IFN-γ, or IFN-γ receptor abolished liver damage in lipopolysaccharide (LPS)/D-galactosa­
mine and Con A-induced hepatitis mice (Hong et al. 2002; Kim et al. 2003) (Figure 9.1).
It is well known that excessive and recurrent liver inflammation and injury can lead to liver
damage, which then progress to fibrosis, cirrhosis, and HCC. It was shown that STAT1 activation
JAK-STAT Signaling in Liver Fibrosis 145

FIGURE 9.1 Role of STATs in promoting liver fibrosis. STAT3 enhances liver fibrosis by inducing both hepatocytes
damage and hepatic stellate cells survival and growth. STAT6 increases the activation of hepatic stellate cells. STAT4
affects the IFN-γ production in natural killer cells, which acts as a pro-fibrotic signal to promote liver injury and fibrosis.

increased cyclin-dependent kinase inhibitor 1 (p21) and interferon regulatory factor (IRF) 1
expression, which inhibited liver regeneration (Sun and Gao 2004; Mair et al. 2011). As well as,
disruption of STAT1 stopped the effect of polyinosinic-polycytidylic acid (poly I:C) treatment on
the liver, which revealed that activation of STAT1-inhibited liver regeneration and promoted the
liver fibrosis progression (Sun et al. 2006; Zhang et al. 2017). Moreover, Zhang et al. showed that
STAT1 suppression plays critical role in the reversion of activated HSCs, which can lead to
reversibility of liver fibrosis (Zhang et al. 2017).
In the hepatocytes, the STAT3 activation inhibits the pro-inflammatory STAT1 signaling and
protects the liver from damage. Interestingly, activation of STAT1 and STAT3 has opposing roles
in liver pathophysiology (inflammation, injury, regeneration, and fibrosis) where STAT1 activation
in the liver is deleterious, whereas STAT3 activation is protective (Gao et al. 2012). Actually,
STAT1 and STAT3 negatively regulate each other by the induction of suppressor of cytokine
signaling-1 (SOCS1) and SOCS3 proteins. In Con A-induced hepatitis model, activated hepatic
STAT1 stimulated Con A-induced hepatitis (Torisu et al. 2008), while activated hepatic STAT3
reduced it (Park et al. 2011). Therefore, the blockage of hepatic STAT1 activation by genetic
modification of some genes prevents the liver injury, while obstruction of hepatic STAT3 increases it
(Hong et al. 2002; Klein et al. 2005; Torisu et al. 2008; Lafdil et al. 2009) (Figure 9.2). Furthermore,
it was reported recently that overexpression of Proline–serine–threonine–phosphatase-interacting
protein2 (PSTPIP2) improves liver inflammation and fibrosis in mice by suppressing the activity of
STAT1 activity, and enhancing the activity of STAT6 activity, which consequently regulates the
macrophage polarization (Yang et al. 2018).
146 JAK-STAT Signaling in Diseases

FIGURE 9.2 Role of STATs in preventing liver fibrosis. STAT3 and STAT5 have hepatoprotective effects by inhibiting
hepatocytes injury and fibrosis. STAT1 has anti-fibrotic effects by inducing apoptosis, cell cycle arrest, and killing of
activated hepatic stellate cells. The STAT1 and STAT4 elevation in natural killer cells contributes to NK cell
cytotoxicity and provides protective effects against viral infection and liver damage.

9.2.2 STAT2
STAT2 is solely activated by the IFN family (IFN-α, IFN-β, and IFN-λ) (Orlent et al. 2011; Kong
et al. 2012b). Albeit STAT2 is well-pronounced in antiviral immune response (Heim and Thimme
2014; Blaszczyk et al. 2015; Shrivastava et al. 2016), its role in liver pathophysiology (liver injury and
fibrosis) is poorly studied. A study performed by Ibrahim et al. (Ibrahim et al. 2017) showed
a significant upregulation of STAT2 and IRF7 in HCV patients with advanced stages of liver fibrosis
when compared to HCV patients with early stages of liver fibrosis. In another study conducted to
characterize the molecular changes associated with hepatic fibrosis at its early stages, HCV patients
showed increased expression of STAT2/IRF9 in their diseased liver when compared to the healthy
livers from controls (Bieche et al. 2005). In conclusion, both of the aforementioned studies suggest
a positive role for STAT2 transcription factor in enhancing liver fibrosis progression.

9.2.3 STAT3
STAT3 plays key roles in the healthy liver; (1) it induces acute-phase response, (2) protects against
hepatocellular damage, and (3) promotes liver regeneration (Wang et al. 2011a; Lee et al. 2016).
However, while many studies suggest the anti-fibrogenic role of STAT3 in liver, some presented
STAT3 as a fibrogenic factor and an excellent target for therapeutic intervention. STAT3
activation in the liver is provoked by a variety of cytokines in a cell type-dependent manner. The
contradiction between different studies in defining the role of STAT3 in liver is mainly attributed
to the nature of the cytokines driving the pathway and the responder cells. In hepatocyte, STAT3
JAK-STAT Signaling in Liver Fibrosis 147

activation is triggered by IL-6, IL-22, and IL-6 family cytokines where it induces an anti-
inflammatory milieu to suppress liver inflammation under most conditions (Kroy et al. 2010;
Wang et al. 2011b; Gao et al. 2012). In kupffer cells, STAT3 is a key downstream signaling
protein of IL-10 (the major anti-inflammatory cytokine) and massive evidence confirms that
STAT3 induces anti-inflammatory signals in myeloid cells, including kupffer cells to control liver
inflammation (Murray 2006). Moreover, the conflicting data on the role of STAT3 in liver fibrosis
is due to the kind of the liver injury models. For example, blockade of STAT3 in hepatocytes
dramatically increased liver injury and inflammation upon chronic CCl4 treatment (Wang et al.
2011b), while it decreased liver inflammation in a model of acute CCl4 administration (Horiguchi
et al. 2010). In T cells, STAT3 might promote or decelerate liver inflammation depending on the
nature of liver injury model being investigated (Figure 9.1). While T cells conditional STAT3­
knockout mice are resistant to liver inflammation induced by Con A injection and show reduced
IL-17 production (Lafdil et al. 2009), inhibition of STAT3 in T cells through SOCS3 overexpression
accelerated acetaminophen hepatotoxicity due to the over production of IFN-γ and tumor necrosis
factor-alpha (TNF-α) (Numata et al. 2007).
Here are some examples for the literature that explored the pro-inflammatory role of STAT3 in
liver. The co-treatment of CCl4-treated Wistar rats with anti-oxidants (Carvedilol and silymarin)
succeeded to restore the normal architecture of the liver and the localization of collagen through
reducing the elevated hepatic IL-6 and STAT-3 levels (Balaha et al. 2016). Primary HSCs isolated
from mouse livers and treated with IL-6 demonstrated upregulation of the reticular fiber
components α-Smooth muscle actin (αSMA) and alpha-1 type I collagen (Col1a) at mRNA and
protein levels, which was associated with increased phosphorylation of STAT3. The fibrogenic
phenotype of HSCs cells was reversed upon using STAT3 synthetic inhibitor (S3I-201) and,
αSMA and Col1a expression levels returned to the normal levels similar to the untreated control
samples (Kagan et al. 2017). Rats treated with diethylnitrosamine (DEN), used to induce
a stepwise histopathological progression in liver (inflammation-fibrosis-HCC), showed a dramatic
up-regulation of JAK2/STAT3 signaling in their liver in a time-dependent manner when compared
to the control group (Ding et al. 2017). Liver tissue sections from cirrhotic and HCC patients
showed high levels of both forms of STAT3 (phosphorylated and non-phosphorylated), with
statistical significant differences between HCC and cirrhosis (Li et al. 2017). Liver injury either
induced by adenovirus infection or CCl4 resulted in the increase of IL-6/IL-12 with the ensuing
activation of STAT3 and STAT4 in naturally occurring regulatory T cells, thus reducing the
suppressive function of regulatory T cells in the liver. The latter is one of the scenarios for
the progression of liver inflammation into fibrosis and autoimmune hepatitis (Chi et al. 2018).
Multiple lines of investigations evidenced that JAK1/STAT3 interact with SMAD pathway to
enhance hepatic fibrosis through the upregulation of TGF-β expression (Ogata et al. 2006; Tang
et al. 2017). S-allyl-cysteine (one of the major antioxidant compounds in aged garlic extract) was
shown to attenuate liver fibrosis in Sprague-Dawley rats treated with CCl4, and that effect was
associated with reduced phosphorylation of SMAD3 and STAT3 and diminished mRNA expression
of TGF-β and IL-6 (Gong et al. 2018). In agreement with this study, cucurbitacin-B treatment in
CCl4 treated mice ameliorated the fibrotic state by diminishing the expression of collagen-1α, αSMA,
TGF-β, and matrix metalloproteinase (MMP) 2. All these effects were associated with the blockade of
STAT3 phosphorylation (Sallam et al. 2018). On the other hand, some studies suggest a protective
role of STAT3 in liver fibrosis. G-protein coupled receptor deficient mice (Gpr110-/- mice) during
CCl4 challenge to induce liver fibrosis displayed lower liver fibrosis; which was attributed to
augmented activation of IL-6/STAT3 pathway as the underlying mechanism (Ma et al. 2017). IL-22
downstream signaling pathway protects liver damage as evident by the finding on the positive
relationship between the level of IL-22BP (an inhibitor of IL-22) and severity of liver fibrosis. The
anti-fibrotic effect of IL-22 is suggested to be through the activation of STAT3 (Khawar et al. 2016).
Kong X et al. reported that IL-22 inhibits liver fibrosis by the induction of HSC senescence, which
was mediated by the activation of STAT3, SOCS3, and p53 pathways (Kong et al. 2012a). Tissue
inhibitor of metalloproteinases-1 (TIMP-1) the well-known player in ECM maintenance that inhibits
the activity of MMPs is one of the downstream gene targets of STAT3 (Sun 2010). TIMP-1 plays
148 JAK-STAT Signaling in Diseases

a protective role against acute and chronic liver injury. Following chronic CCl4 treatment, the levels of
the circulating and liver specific TIMP-1 were markedly diminished in conditional hepatocyte-specific
STAT3 knockout mice (Wang et al. 2011c; Kasembeli et al. 2018).

9.2.4 STAT4
STAT4 is a key transcription factor in the process of Th1 cell differentiation and in promoting
cellular-mediated immunity (Wang et al. 2015). STAT4 is expressed mainly in myeloid cells (Frucht
et al. 2000), thymus, and testes and to a very low level in resting human T cells. Signaling through
STAT4 is induced by arrays of cytokines, including IL-12 (Bacon et al. 1995; Wang et al. 2015), IL-2
(Wang et al. 1999), IFN-α/β (Cho et al. 1996; Miyagi et al. 2007), and IL-17 (Subramaniam et al.
1999). The activated STAT4 regulates the transcription of myriad genes. The most prominent STAT4­
target gene is IFN-γ. IL-12-dependent STAT4 activation is well studied in liver diseases. IL-12/STAT4
signaling is a key regulator of tissue inflammation, fibrogenesis, and antiviral defense (Edlich et al.
2012; Kong et al. 2012b; Wang et al. 2015). In a mouse model of Schistosoma infection, the animals
treated with anti-IL-12 had granuloma with larger size and intensified fibrosis (Cheng et al. 2012),
which suggests a protective role of IL-12/STAT4 axis against liver fibrosis in these animals. In a mouse
model of NASH, the knockdown of protein inhibitor of activated STAT4 (PIAS4) using short hairpin
RNA (shRNA) dramatically attenuated liver fibrosis induced by high-fat high-carbohydrate (HFHC)
diet through the downregulation of a panel of pro-fibrogenic ECM-related genes, including collagens
(type I and type III), smooth muscle actin, and tissue inhibitors of metalloproteinase, the anti-fibrotic
effect of PIAS4 silencing is mediated by the blockade of SMAD3-(a potent pro-fibrogenic transcrip­
tion factor) related signal pathway (Xu et al. 2016).
The genetic variation in STAT4 may expose the host to the risk of fibrosis; confirming the
indispensable role of this transcription factor in the liver health. Examples for the latter, genotyping
of STAT4 single nucleotide polymorphism (SNP) (rs7574865) in 160 liver-transplanted patients with
HCV recurrence revealed that STAT4-T-allele is highly linked to the development of advanced
fibrosis (Eurich et al. 2013). For the same STAT4 SNP (rs7574865), Jiang et al. correlated G-allele
with the risk of HBV-related liver cirrhosis (Jiang et al. 2015). In another study, the author
elaborated the impact of STAT4 genetic variation on the outcome of STAT4 signaling with the
ensuing consequence on the liver pathology. In this study, the Caucasian HBV-patients who are
carriers of STAT4 rs7574865 GG genotype expressed at very low levels STAT4 in liver, peripheral
blood mononuclear cells (PBMCs), and in NK cells. Upon stimulating NK cells (isolated from the
patients with the risk allele) with IL-12/IL-18, the cells showed impaired STAT4 phosphorylation
associated with a reduction in IFN-γ secretion. NK cells malfunction due to the impaired STAT4
phosphorylation-dependent IFN-γ production likely contributes to hepatic inflammation and
fibrosis in HBV patients (El Sharkawy et al. 2018). On the other hand, some studies suggest
a fibrotic role of STAT4 in the liver. Wang et al. used IL-35 gene-modified mesenchymal stem cells
(IL-35-MSCs) to deliver IL-35 gene and investigated their protective effects in Con A-induced
autoimmune hepatitis. The study showed the ability of these genetically modified cells to migrate to
the diseased liver where it narrowed the necrosis area. The curative effect of these cells was
attributed to their ability to interfere with JAK1-STAT1/STAT4 signaling cascade with the
consequent decrease in IFN-γ production secreted by liver mononuclear cells (Wang et al. 2018).
IL-12 deficiency rendered the mice to resist Con A-induced T cell hepatitis (Zhu et al. 2007),
whereas IL-12 overexpression in the liver induces liver injury (Kong et al. 2012b). Similarly, IL-12
treatment has been shown to induce liver inflammation in multiple animal models (Harada et al.
2004; Chang et al. 2007), which is likely mediated by the subsequent production of IFN-γ in NK
cells and natural killer T-cells (NKT) (Subleski et al. 2006). This suggests that IL-12/STAT4 acts as
a pro-inflammatory signal pathway to promote liver injury and fibrosis (Figure 9.1). Collectively,
STAT4-IL-12 dependent activation in immune cells has double sword edges, it may induce
inflammation in the liver, which is likely to promote to liver injury and fibrosis, but may also
protect against infection and hence, ameliorating liver injury and fibrosis.
JAK-STAT Signaling in Liver Fibrosis 149

9.2.5 STAT5
The STAT5a and STAT5b proteins are encoded by two different genes located on chromosome 17
of human and chromosome 11 of mice. The STAT5 proteins are activated mainly by growth
hormone (GH) and interleukins (IL-2, IL-3, IL-5), as well as some other cytokines (Mair et al.
2011). In hepatocytes, the growth hormone activates STAT5, which controls the expression of
many hepatic genes essential for growth, metabolism, and differentiation in the liver (Gao 2005).
Additionally, STAT5 regulates many hepatoprotective factors that inhibit hepatocyte cell death.
The GH/STAT5 signaling pathway regulates the expression of insulin-like growth factor 1 (IGF-1)
and epidermal growth factor receptor (EGFR) in hepatocytes and muscle cells (Woelfle et al.
2003; Mair et al. 2011). It was reported that high GH and low IGF-1 serum levels are associated
with human liver cirrhosis and the level of IGF-1 correlates with liver disease progression. IGF-1
low serum levels lead to hepatocellular dysfunction (Donaghy et al. 1995; Caregaro et al. 1998). It
was demonstrated that loss of STAT5 or GH diminished the liver regeneration after partial
hepatectomy (PHx). Moreover, STAT5 loss caused hepatosteatosis defective GH/STAT5/IGF-1
signaling, insulin resistance, and increased the activity of STAT1 (Cui et al. 2007).
In a cholestasis mouse model, IGF-1 is an important target of GH/STAT5 signals and the GH/
STAT5/IGF-1 axis protects liver from injury (Sobrevals et al. 2010). It has been demonstrated
that GH/STAT5 regulates the expression of hepatocyte nuclear factor 6 (HNF6) which stimulates
liver regeneration (Wang et al. 2008). Also, STAT5 upregulates the expression of prolactin
receptor (PRLP) and leukemia inhibitory factor receptor (LIFR), which play crucial role in liver
regeneration and protection (Simon-Holtorf et al. 2006; Blaas et al. 2010). Deletion of STAT5 in
mice caused GH resistance, and disruption of IGF-1 transcription induced by GH, which leads to
cell damage and apoptosis (Blaas et al. 2010). Also, it has been shown that administration of
IGF-1 improved the liver fibrosis by increasing the mitochondrial function, reducing the oxidative
liver damage, and increasing the MMPs expression (Sobrevals et al. 2010). Consistently, IGF-
1-stimulates HSCs to produce hepatocyte growth factor (HGF), which promotes HSCs apoptosis
(Kim et al. 2005).
During liver inflammation or injury, hepatocytes secrete excessive amount of TGF-β, which
induces extracellular matrix accumulation and ultimately leads to the development of liver fibrosis
(Canbay et al. 2004; Jeong et al. 2006). STAT5 provides protection against fatty liver disease and
liver injury (Mueller et al. 2011; Friedbichler et al. 2012; Baik et al. 2017) In cholestasis mouse
model, it was confirmed that STAT5 has an anti-fibrotic effect as it protects the liver from injury
and fibrosis (Kong et al. 2012b). STAT5 prevents the liver fibrosis induced by the deletion of
multidrug resistance gene 2. Actually, the deletion of both STAT5 (hepatocyte and cholangiocyte)
and multidrug resistance gene 2 leads to severe liver fibrosis and increases the number of
apoptotic hepatocytes. Moreover, these deletions are associated with reduced expression of
essential hepatoprotective genes such as HNF6, EGFR, PRLP, and LIFR (Blaas et al. 2010).
Hosui et al. discovered a relationship between STAT3, STAT5, and TGF-β in the pathogenesis of
liver fibrosis. In STAT5-null mouse embryonic fibroblasts (MEF) cells, it was reported that TGF-β
levels were highly elevated and treatment with STAT5 caused the suppression of TGF-β. Also, the
STAT5 loss in hepatocytes after CCl4 treatment enhanced the TGF-β levels and GH-induced
STAT3 activity. Therefore, it was proposed that STAT5 prevents liver fibrogenes by downregulating
TGF-β and STAT3 (Figure 9.2). Nevertheless, it was reported that the deletion of STAT5 increases
the sensitivity of HSC or Kupffer cells to TGF-β and promoted liver fibrosis (Hosui et al. 2009).
Furthermore, hepatic STAT5 deletion promoted CCl4-induced HCC in wild-type mice (Hosui et al.
2009) and induced the development of HCC in hepatic glucocorticoid receptor knockout mice
(Mueller et al. 2011) or in growth hormone transgenic mice (Friedbichler et al. 2012).
Several studies documented the role of STAT5 in tumorigenesis (Moriggl et al. 1999; Haddad
et al. 2013). Interestingly, the role of STAT 5 in liver carcinogenesis is controversial. It was reported
that STAT5 activation causes cellular transformation by stimulating progression of cell cycle and
suppression of cell apoptosis. Activated STAT5 upregulates the expression of cell proliferation,
invasion, metastasis, and anti-apoptotic-related genes, which promotes tumorigenesis (Ferbeyre and
150 JAK-STAT Signaling in Diseases

Moriggl 2011). However, other studies proved that STAT5 has antioncogenic function in liver
cancer as hepatic deletion of STAT5 led to HCC development after CCl4 treatment. Additionally,
the loss of hepatic STAT5 is compensated by STAT3 upregulation, which stimulates hepatic
tumorigenesis (Hosui et al. 2009). The activation of STAT5 induces the expression of cell cycle
inhibitor (p15INK4B) in hepatocytes, which has a protective effect against liver carcinogenesis (Yu
et al. 2011). Also, the hepatic STAT5 deletion enhanced liver cancer caused by hyper-activated
growth hormone signaling (Friedbichler et al. 2012). Therefore, it was suggested that STAT5 has
hepato-protective and anti-steatogenic effects and can be a tumor suppressor in liver carcinogensis
(Mueller et al. 2011).

9.2.6 STAT6
STAT6 has both a pro and anti-inflammatory effect. Activation on of STAT6 in the liver is
induced by Th1 cytokine (IL-12) and Th2 cytokines mainly (IL-4, IL-13) (Wurster et al. 2000;
Jaruga et al. 2003). STAT6 has important roles in the differentiation of Th2 lymphocyte (Kaplan
et al. 1996; Murphy and Reiner 2002). Moreover, STAT6 plays key roles in controlling liver injury,
inflammation, and fibrosis. The functions of STAT6 have been studied in different liver injury
models including Con A-injection, ischemia/reperfusion, and LPS-treatment (Kato et al. 2000;
Lentsch et al. 2001; Iff et al. 2009).
In HSCs, the STAT6 is predominately induced by both IL-4 and IL-13 and the blockage of
STAT6 using small interfering RNA (siRNA) attenuates the activation of HSC in vitro (Aoudje­
hane et al. 2008). The IL-4 and IL-13 roles in liver fibrosis have been extensively examined and
documented (Barron and Wynn 2011; Du et al. 2016; Gieseck et al. 2016), while the STAT6 roles
in liver inflammation and fibrosis is controversial and needs comprehensive studies to be largely
clear. Studies on different liver injury models suggested that STAT6 have either a harmful or
protective effect on liver fibrogenesis.
Previously, it was shown that STAT6-deficient mice with Schistosoma mansoni infection have
reduced collagen deposition in the liver and smaller granulomas than wild type mice (Kaplan
et al. 1996). However, Liu et al. (2011) has confirmed that ERK1/2 pathway rather than STAT6 in
HSCs contributes to IL-13 induction of connective tissue growth factor (CTGF), which stimulates
deposition of extracellular matrix in the liver. IL-13 has been proved to promote liver fibrogenesis
in Schistosoma mansoni infection model and the blockage of IL-13 reduces the liver fibrosis. The
IL-13 upregulates several fibrotic proteins in HSCs and induces the HSCs activation (Fallon et al.
2000; Chiaramonte et al. 2001; Du et al. 2016).
Recently, it has been shown that IL-4 induces the activation of STAT6 in a different liver injury
models (Njoku et al. 2009). IL-4 has pro-inflammatory and pro-fibrotic effects (Weng et al. 2018).
IL-4 stimulates the activation of HSCs and collagens production (Aoudjehane et al. 2008; Jin
et al. 2011). Upregulation of IL-4 expression has been detected is in the fibrotic liver of baboons
infected with Schistosoma mansoni, and the liver fibrosis in Schistosoma mansoni-infected mice
was significantly reduced by the blockage of IL-4 (Farah et al. 2000). The role of IL-4/STAT6 in
hepatitis was confirmed in SOCS-1 deficiency-induced hepatitis (Naka et al. 2001). Moreover, the IL­
4/STAT6 signaling has important functions in T-cell mediated hepatitis induced by Con A–treatment.
It was suggested that IL-4/STAT6 enhances the eotaxins expression in sinusoidal endothelial cells and
hepatocytes, stimulates IL-5 expression, and induces eosinophils and neutrophils infiltration into the
liver and which leads to hepatitis. Also, it was documented that the activation of STAT6 in
IL-4-deficient mice was significantly suppressed after Con A injection (Kaneko et al. 2000;
Jaruga et al. 2003).
On the contrary, Ryan et al. recently showed that IL-4 deficient mice have increased liver injury
and inflammation after acetaminophen injection (Ryan et al. 2012). Moreover, the severity of liver
injury was reduced by the administrating of recombinant IL-4. It was proved in acetaminophen-
induced injury model that IL-4 has hepatoprotective function by upregulating the synthesis of
hepatic glutathione (GSH). In addition, several studies showed activation of STAT6 by both IL-4
JAK-STAT Signaling in Liver Fibrosis 151

and IL-13 protected the liver from ischemia/reperfusion injury. Moreover, STAT6 deficient mice
were highly susceptible to endotoxin-induced liver injury (Lentsch et al. 2001). STAT6 has
potential anti-inflammatory effect as it inhibited liver inflammation and protected endothelial
cell and hepatocyte against damage (Kato et al. 2000) Concisely, the STAT6 activation promoted
liver injury and fibrogenesis in T cell hepatitis model while it protected against ischemia/reperfusion
and drug-induced liver injury (Gao et al. 2012).

9.3 Conclusion
Liver fibrosis is severe hepatocytes damage, which leads to end-stage liver disease. The chronic
inflammation, continuous liver injury, and inadequate recovery from hepatocyte damage trigger
numerous cytokines and cellular events that lead to the liver injury, the collagen deposition, and
the interruption of the liver normal architecture. Several studies have reported the crucial role of
JAK-STAT signaling pathway in hepatic fibrogenesis.
The current chapter summarizes the biological significance of different STATs signaling proteins
in controlling liver fibrosis as shown in Figures 9.1 and 9.2. STAT1 has anti-fibrotic effect via
inducing HSC apoptosis and cell cycle arrest or has pro-inflammatory signaling, which stimulates
liver fibrosis. Interestingly, STAT1 and STAT3 negatively regulate each other as the activation of
STAT3 prevents the deleterious effect of STAT1. STAT2 has proven antiviral effect but its role in
liver injury and fibrosis is not well determined and needs to be further studied. STAT3 may have
a pro-inflammatory or anti-fibrogenic role in liver. STAT3 induces hepatic stellate cells survival
and growth, which promotes liver fibrosis; whereas STAT3, in hepatocytes, have anti-fibrotic and
hepatoprotective effects via preventing the liver injury and fibrosis (Figure 9.1). STAT4 is very
important for cellular-mediated immunity. Activation of STAT4 may induce inflammation and
liver injury or may also protect against infection and hence, preventing liver damage and fibrosis.
STAT5 controls many hepatoprotective factors that prevent hepatocyte cell death and increased
liver regeneration. Moreover, STAT5 provides protection against fatty liver disease, liver injury,
fibrosis, and hepatocellular carcinoma. STAT6 has both a pro and anti-inflammatory effect.
However, STAT6 roles in liver inflammation and fibrosis are controversial and need comprehensive
studies to be largely clear.

REFERENCES
Akçalı, K., and M. M. Aydın. 2018. Liver Fibrosis. Turk J Gastroenterol 29:8.

Aoudjehane, L., A. Pissaia, O. Scatton, P. Podevin, P.P. Massault, S. Chouzenoux, O. Soubrane,

Y. Calmus, and F. Conti. 2008. Interleukin-4 Induces the Activation and Collagen Production of
Cultured Human Intrahepatic Fibroblasts via the STAT-6 Pathway. Lab Investig 9:973–85.
Bacon, C. M., E. F. Petricoin, 3rd, J. R. Ortaldo, R. C. Rees, A. C. Larner, J. A. Johnston, and
J. J. O’Shea. 1995. Interleukin 12 Induces Tyrosine Phosphorylation and Activation of STAT4 in
Human Lymphocytes. Proc Natl Acad Sci USA 92(16):7307–11.
Baik, M., J. Kim, M. Y. Piao, H. J. Kang, S. J. Park, S.W. Na, S. H. Ahn, and J. H. Lee. 2017. Deletion of
Liver-Specific STAT5 Gene Alters the Expression of Bile Acid Metabolism Genes and Reduces
Liver Damage in Lithogenic Diet-Fed Mice. J Nutr Biochem 39:59–67.
Balaha, M., S. Kandeel, and W. Barakat. 2016. Carvedilol Suppresses Circulating and Hepatic IL-6
Responsible for Hepatocarcinogenesis of Chronically Damaged Liver in Rats. Toxicol Appl
Pharmacol 311:1–11.
Barron, L., and T. A. Wynn. 2011. Fibrosis Is Regulated by Th2 and Th17 Responses and by Dynamic
Interactions between Fibroblasts and Macrophages. Am J Physiol Gastrointest Liver Physiol 300(5):
G723–8.
Bieche, I., T. Asselah, I. Laurendeau, D. Vidaud, C. Degot, V. Paradis, P. Bedossa, D. C. Valla,
P. Marcellin, and M. Vidaud. 2005. Molecular Profiling of Early Stage Liver Fibrosis in Patients
with Chronic Hepatitis C Virus Infection. Virology 332(1):130–44.
152 JAK-STAT Signaling in Diseases

Blaas, L., J. Kornfeld, D. Schramek, M. Musteanu, J. Gumhold, F. Van Zijl, D. Schneller, and
H. Esterbauer. 2010. Europe PMC Funders Group Disruption of the Growth Hormone — Signal
Transducer and Activator of Transcription 5 — Insulin like Growth Factor 1 Axis Severely
Aggravates Liver Fibrosis in a Mouse Model of Cholestasis. Hepatology 51(4):1319–26.
Blaszczyk, K., A. Olejnik, H. Nowicka, L. Ozgyin, Y. L. Chen, S. Chmielewski, K. Kostyrko, J. Wesoly,
B. L. Balint, C. K. Lee, and H. A. Bluyssen. 2015. STAT2/IRF9 Directs a Prolonged ISGF3-like
Transcriptional Response and Antiviral Activity in the Absence of STAT1. Biochem J 466(3):511–24.
Brenner, D. A. 2013. Reversibility of Liver Fibrosis. Gastroenterol Hepatol (NY) 9(11):737–39.
Canbay, A., S. Friedman, and G. J. Gores. 2004. Apoptosis: The Nexus of Liver Injury and Fibrosis.
Hepatology 39(2):273–78.
Caregaro, L., F. Alberino, P. Angeli, and A. Gatta. 1998. Insulin-like Growth Factor 1 (IGF-1) in Liver
Cirrhosis: A Marker of Hepatocellular Dysfunction? J Hepatol 29(2):342.
Chang, C. J., Y. H. Chen, K. W. Huang, H. W. Cheng, S. F. Chan, K. F. Tai, and L. H. Hwang. 2007.
Combined GM-CSF and IL-12 gene Therapy Synergistically Suppresses the Growth of Orthotopic
Liver Tumors. Hepatology 45(3):746–54.
Cheng, Y. L., W. J. Song, W. Q. Liu, J. H. Lei, Z. Kong, and Y. L. Li. 2012. The Effects of Interleukin
(IL)-12 and IL-4 Deficiency on Worm Development and Granuloma Formation in Schistosoma
Japonicum-Infected Mice. Parasitol Res 110(1):287–93.
Chi, G., X. X. Feng, Y. X. Ru, T. Xiong, Y. Gao, H. Wang, Z. L. Luo, R. Mo, F. Guo, Y. P. He,
G. M. Zhang, D. A. Tian, and Z. H. Feng. 2018. TLR2/4 Ligand-Amplified Liver Inflammation
Promotes Initiation of Autoimmune Hepatitis Due to Sustained IL-6/IL-12/IL-4/IL-25 Expression.
Mol Immunol 99:171–81.
Chiaramonte, M. G., A.W. Cheever, J. D. Malley, D. D. Donaldson, and T. A. Wynn. 2001. Studies of
Murine Schistosomiasis Reveal Interleukin-13 Blockade as a Treatment for Established and
Progressive Liver Fibrosis. Hepatology 34(2):273–82.
Cho, S. S., C. M. Bacon, C. Sudarshan, R. C. Rees, D. Finbloom, R. Pine, and J. J. O’Shea. 1996.
Activation of STAT4 by IL-12 and IFN-Alpha: Evidence for the Involvement of Ligand-induced
Tyrosine and Serine Phosphorylation. J Immunol 157(11):4781–89.
Cui, Y., A. Hosui, R. Sun, K. Shen, O. Gavrilova, W. Chen, M.C. Cam, B. Gao, G.W. Robinson, and
L. Hennighausen. 2007. Loss of Signal Transducer and Activator of Transcription 5 Leads to
Hepatosteatosis and Impaired Liver Regeneration. Hepatology 46(2):504–13.
Darnell, J.E., Jr., I.M. Kerr, and G. R. Stark. 1994. JAK–STAT Pathways and Transcriptional Activation
in Response to IFNs and Other Extracellular Signaling Proteins. Science 264(5164):1415–21.
Ding, Y. F., Z. H. Wu, Y. J. Wei, L. Shu, and Y. R. Peng. 2017. Hepatic Inflammation-fibrosis-cancer Axis
in the Rat Hepatocellular Carcinoma Induced by Diethylnitrosamine. J Cancer Res Clin Oncol 143
(5):821–34.
Donaghy, A., R. Ross, A. Gimson, S.C. Hughes, J. Holly, and R. Williams. 1995. Growth Hormone,
Insulinlike Growth Factor-1, and Insulinlike Growth Factor Binding Proteins 1 and 3 in Chronic
Liver Disease. Hepatology 21(3):680–88.
Donnelly, R., H. Dickensheets, and T. O’Brien. 2011. Interferon-Lambda and Therapy for Chronic
Hepatitis C Virus Infection. Trends Immunol 32(9):443–50.
Du, P., Q. Ma, Z. De Zhu, G. Li, Y. Wang, Q. Q. Li, Y. F. Chen, Z. Z. Shang, J. Zhang, and L. Zhao. 2016.
Mechanism of Corilagin Interference with IL-13/STAT6 Signaling Pathways in Hepatic Alternative
Activation Macrophages in Schistosomiasis-Induced Liver Fibrosis in Mouse Model. Eur
J Pharmacol 793:119–126.
Edlich, B., G. Ahlenstiel, A.Z. Azpiroz, J. Stoltzfus, M. Noureddin, E. Serti, J. J. Feld, T. J. Liang, Y. Rotman,
and B. Rehermann. 2012. Early Changes in Interferon Signaling Define Natural Killer Cell Response
and Refractoriness to Interferon-Based Therapy of Hepatitis C Patients. Hepatology 55(1):39–48.
El Sharkawy, R., K. Thabet, P. Lampertico, S. Petta, A. Mangia, T. Berg, M. Metwally, A. Bayoumi,
A. Boonstra, W. P. Brouwer, A. Smedile, M. L. Abate, A. Loglio, M. W. Douglas, A. Khan, R. Santoro,
J. Fischer, D. J. Leeming, C. Liddle, J. George, and M. Eslam. 2018. A STAT4 Variant Increases Liver
Fibrosis Risk in Caucasian Patients with Chronic Hepatitis B. Aliment Pharmacol Ther 48(5):564–73.
Eurich, D., S. Boas-Knoop, B. Struecker, R. Neuhaus, P. Neuhaus, and M. Bahra. 2013. Genetic Variants
of STAT-4 Affect the Development of Graft Fibrosis after Liver Transplantation for HCV-induced
Liver Disease. Transplantation 95(1):203–08.
JAK-STAT Signaling in Liver Fibrosis 153

Fallon, P. G., E. J. Richardson, G. J. McKenzie, and A. N. J. McKenzie. 2000. Schistosome Infection of


Transgenic Mice Defines Distinct and Contrasting Pathogenic Roles for IL-4 and IL-13: IL-13 Is
a Profibrotic Agent. J Immunol 164(5):2585–91.
Farah, I. O., P. W. Mola, T. M. Kariuki, M. Nyindo, R. E. Blanton, and C. L. King. 2000. Repeated
Exposure Induces Periportal Fibrosis in Schistosoma Mansoni-Infected Baboons: Role of TGF-
and IL-4. J Immunol 164(10):5337–43.
Ferbeyre, G., and R. Moriggl. 2011. The Role of Stat5 Transcription Factors as Tumor Suppressors or
Oncogenes. Biochim Biophys Acta Rev Cancer 1815(1):104–14.
Friedbichler, K., M. Themanns, K. M. Mueller, M. Schlederer, J. Kornfeld, L. M. Terracciano,
A. V. Kozlov, S. Haindl, L. Kenner, T. Kolbe, M. Mueller, K. J. Snibson, M. H. Heim, and
R. Moriggl. 2012. Growth-Hormone–Induced Signal Transducer and Activator of Transcription 5
Signaling Causes Gigantism, Inflammation, and Premature Death but Protects Mice From Aggres-
sive Liver Cancer. Hepatology 55(3):941–52.
Frucht, D. M., M. Aringer, J. Galon, C. Danning, M. Brown, S. Fan, M. Centola, C. Y. Wu, N. Yamada,
H. El Gabalawy, and J. J. O’Shea. 2000. Stat4 is Expressed in Activated Peripheral Blood
Monocytes, Dendritic Cells, and Macrophages at Sites of Th1-mediated Inflammation. J Immunol
164(9):4659–64.
Gao, B. 2005. Cytokines, STATs and Liver Disease. Cell Mol Immunol 2(2):92–100.
Gao, B., H. Wang, F. Lafdil, and D. Feng. 2012. STAT Proteins – Key Regulators of Anti-Viral
Responses, Inflammation, and Tumorigenesis in the Liver. J Hepatol 57(2):430–41.
Ghoreschi, K., A. Laurence, and J. J. O’Shea. 2009. Janus Kinases in Immune Cell Signaling. Immunol
Rev 228(1):273–87.
Gieseck, R. L., T. R. Ramalingam, K. M. Hart, K. M. Vannella, D. A. Cantu, W. Lu, S. Ferreira-González,
S. J. Forbes, and L. Vallier. 2016. Interleukin-13 Activates Distinct Cellular Pathways Leading to
Ductular Reaction, Steatosis, and Fibrosis. Immunity 45(3):145–58.
Gong, Z., H. Ye, Y. Huo, L. Wang, Y. Huang, M. Huang, and X. Yuan. 2018. S-Allyl-Cysteine Attenuates
Carbon Tetrachloride-Induced Liver Fibrosis in Rats by Targeting STAT3/SMAD3 Pathway. Am
J Transl Res 10(5):1337–46.
Haddad, B. R., L. Gu, T. Mirtti, A. Dagvadorj, P. Vogiatzi, D.T. Hoang, R. Bajaj, B. Leiby, E. Ellsworth,
S. Blackmon, C. Ruiz, M. Curtis, P. Fortina, A. Ertel, C. Liu, H. Rui, T. Visakorpi, L. Bubendorf,
C. D. Lallas, E. J. Trabulsi, P. McCue, L. Gomella, and M. T. Nevalainen. 2013. STAT5A/B Gene
Locus Undergoes Amplification during Human Prostate Cancer Progression. Am J Pathol 182
(6):2264–75.
Harada, N., M. Shimada, S. Okano, T. Suehiro, Y. Soejima, Y. Tomita, and Y. Maehara. 2004. IL-12
Gene Therapy is An Effective Therapeutic Strategy for Hepatocellular Carcinoma in Immunosup-
pressed Mice. J Immunol 173(11):6635–44.
Heim, M. H., and R. Thimme. 2014. Innate and Adaptive Immune Responses in HCV Infections.
J Hepatol 61(1 Suppl):S14–25.
Hong, F., B. Jaruga, W. H. Kim, S. Radaeva, O. N. El-Assal, Z. Tian, V. A. Nguyen, and B. Gao. 2002.
Opposing Roles of STAT1 and STAT3 in T Cell-Mediated Hepatitis: Regulation by SOCS. J Clin
Invest 110(10):1503–13.
Horiguchi, N., F. Lafdil, A. M. Miller, O. Park, H. Wang, M. Rajesh, P. Mukhopadhyay, X. Y. Fu,
P. Pacher, and B. Gao. 2010. Dissociation between Liver Inflammation and Hepatocellular Damage
Induced by Carbon Tetrachloride in Myeloid Cell-Specific Signal Transducer and Activator of
Transcription 3 Gene Knockout Mice. Hepatology 51(5):1724–34.
Hosui, A., A. Kimura, D. Yamaji, B. M. Zhu, R. Na, and L. Hennighausen. 2009. “Loss of STAT5
Causes Liver Fibrosis and Cancer Development through Increased TGF-β and STAT3 Activation.”
J Exp Med 206(4):819–31. doi: 10.1084/jem.20080003.
Ibrahim, M. K., A. Khedr, N. G. Bader El Din, A. Khairy, and M. K. El Awady. 2017. Increased
Incidence of Cytomegalovirus Coinfection in HCV-Infected Patients with Late Liver Fibrosis is
Associated with Dysregulation of JAK–STAT Pathway. Sci Rep 7(1):10364.
Iff, J., W. Wang, T. Sajic, N. Oudry, E. Gueneau, G. Hopfgartner, E. Varesio, and I. Szanto. 2009.
Differential Proteomic Analysis of STAT6 Knockout Mice Reveals New Regulatory Function in
Liver Lipid Homeostasis. J Proteome Res 8(10):4511–24.
154 JAK-STAT Signaling in Diseases

Inagaki, Y., T. Nemoto, Z. Miwa Kushida, Y. Sheng, K. Higa, K. Ikeda, F. Shirasaki, K. Takehara,
K. Sugiyama, M. Fujii, H. Yamauchi, A. Nakao, B. de Crombrugghe, T. Watanabe, and I. Okazaki.
2003. Interferon Alfa Down-Regulates Collagen Gene Transcription and Suppresses Experimental
Hepatic Fibrosis in Mice. Hepatology 38(4):890–99.
Jaruga, B., F. Hong, R. Sun, S. Radaeva, and B. Gao. 2003. Crucial Role of IL-4/STAT6 in T
Cell-Mediated Hepatitis: Up-Regulating Eotaxins and IL-5 and Recruiting Leukocytes. J Immunol
171(6):3233–44.
Jeong, W. I., O. Park, S. Radaeva, and B. Gao. 2006. STAT1 Inhibits Liver Fibrosis in Mice by Inhibiting

Stellate Cell Proliferation and Stimulating NK Cell Cytotoxicity. Hepatology 44(6):1441–51.

Jiang, D. K., X. P. Ma, X. Wu, L. Peng, J. Yin, Y. Dan, H. X. Huang, D. L. Ding, L. Y. Zhang, Z. Shi,

P. Zhang, H. Yu, J. Sun, S. Lilly Zheng, G. Deng, J. Xu, Y. Liu, J. Guo, G. Cao, and L. Yu. 2015.
Genetic Variations in STAT4,C2,HLA-DRB1 and HLA-DQ Associated with Risk of Hepatitis B
Virus-related Liver Cirrhosis. Sci Rep 5:16278.
Jin, Z., R. Sun, H. Wei, X. Gao, Y. Chen, and Z. Tian. 2011. Accelerated Liver Fibrosis in Hepatitis
B Virus Transgenic Mice: Involvement of Natural Killer T Cells. Hepatology 53(1):219–29.
Kagan, P., M. Sultan, I. Tachlytski, M. Safran, and Z. Ben-Ari. 2017. Both MAPK and STAT3 Signal
Transduction Pathways are Necessary for IL-6-Dependent Hepatic Stellate Cells Activation. PLoS
One 12(5):e0176173.
Kaneko, B.Y., M. Harada, T. Kawano, M. Yamashita, Y. Shibata, F. Gejyo, T. Nakayama, and
M. Taniguchi. 2000. Augmentation of Valpha14 NKT Cell – Mediated Cytotoxicity by Interleukin
4 in an Autocrine Mechanism Resulting in the Development of Concanavalin A – Induced
Hepatitis. J Exp Med 191(1):105–14.
Kaplan, M. H., U. Schindler, S. T. Smiley, and M. J. Grusby. 1996. Stat6 Is Required for Mediating
Responses to IL-4 and for the Development of Th2 Cells. Immunity 4(3):313–19.
Kasembeli, M. M., U. Bharadwaj, P. Robinson, and D. J. Tweardy. 2018. Contribution of STAT3 to
Inflammatory and Fibrotic Diseases and Prospects for its Targeting for Treatment. Int J Mol Sci 19
(8):pii: E2299.
Kato, A., H. Yoshidome, M. J. Edwards, and A. B. Lentsch. 2000. Reduced Hepatic Ischemia/Reperfu­
sion Injury by IL-4: Potential Anti- Inflammatory Role of STAT6. Inflamm Res 49(6):275–79.
Khawar, M. B., F. Azam, N. Sheikh, and K. Abdul Mujeeb. 2016. How Does Interleukin-22 Mediate
Liver Regeneration and Prevent Injury and Fibrosis? J Immunol Res 2016:2148129.
Kim, W., F. Hong, S. Radaeva, B. Jaruga, S. Fan, and B. Gao. 2003. STAT1 Plays an Essential Role in
LPS/D-Galactosamine-Induced Liver Apoptosis and Injury. Am J Physiol Gastrointest Liver
Physiol 20892:761–68.
Kim, W. H., K. Matsumoto, K. Bessho, and T. Nakamura. 2005. Growth Inhibition and Apoptosis in
Liver Myofibroblasts Promoted by Hepatocyte Growth Factor Leads to Resolution from Liver
Cirrhosis. Am J Pathol 166(4):1017–28.
Kisseleva, T., and D. A. Brenner. 2008. Mechanisms of Fibrogenesis. Exp Biol Med (Maywood) 233
(2):109–22.
Klein, C., T. Wüstefeld, U. Assmus, T. Roskams, S. Rose-John, M. Müller, M. P. Manns, M. Ernst, and
C. Trautwein. 2005. Triggers Liver Protection in T Cell – Mediated Liver Injury Find the Latest
Version: The IL-6–Gp130–STAT3 Pathway in Hepatocytes Triggers Liver Protection in T Cell –
Mediated Liver Injury. J Clin Invest 115(4):860–69.
Kong, X., D. Feng, H. Wang, F. Hong, A. Bertola, F. S. Wang, and B. Gao. 2012a. Interleukin-22
Induces Hepatic Stellate Cell Senescence and Restricts Liver Fibrosis in Mice. Hepatology 56
(3):1150–59.
Kong, X., N. Horiguchi, M. Mori, and B. Gao. 2012b. Cytokines and STATs in Liver Fibrosis. Front
Physiol 3:1–7.
Kroy, D. C., N. Beraza, D. F. Tschaharganeh, L. E. Sander, S. Erschfeld, A. Giebeler, C. Liedtke,
H. E. Wasmuth, C. Trautwein, and K. L. Streetz. 2010. Lack of Interleukin-6/glycoprotein 130/
signal Transducers and Activators of Transcription-3 Signaling in Hepatocytes Predisposes to Liver
Steatosis and Injury in Mice. Hepatology 51(2):463–73.
Lafdil, F., H. Wang, O. Park, W. Zhang, Y. Moritoki, S. Yin, X. Y. Fu, M. E. Gershwin, Z. X. Lian, and
B. Gao. 2009. Myeloid STAT3 Inhibits T cell-mediated Hepatitis by Regulating T Helper 1
Cytokine and Interleukin-17 Production. Gastroenterology 137(6):2125–35 e1–2.
JAK-STAT Signaling in Liver Fibrosis 155

Lee, S. C., H. J. Jeong, S. K. Lee, and S. J. Kim. 2016. Hypoxic Conditioned Medium From Human
Adipose-Derived Stem Cells Promotes Mouse Liver Regeneration through JAK/STAT3 Signaling.
Stem Cells Transl Med 5(6):816–25.
Lentsch, A. B., A. Kato, B. Davis, W. Wang, C. Chao, and M. J. Edwards. 2001. STAT4 and STAT6
Regulate Systemic Inflammation and Protect against Lethal Endotoxemia. J Clin Invest 108
(10):1475–82.
Li, Y. P., W. Z. Wang, X. Q. Chen, L. B. Li, Z. Y. Liang, K. Ru, and J. N. Li. 2017. Signal Transducer and
Activator of Transcription 3 for the Differentiation of Hepatocellular Carcinoma from Cirrhosis.
Chin Med J (Engl) 130(22):2686–90.
Ma, B., J. Zhu, J. Tan, Y. Mao, L. Tang, C. Shen, H. Zhang, Y. Kuang, J. Fei, X. Yang, and Z. Wang.
2017. Gpr110 Deficiency Decelerates Carcinogen-induced Hepatocarcinogenesis via Activation of
the IL-6/STAT3 Pathway. Am J Cancer Res 7(3):433–47.
Machida, K., H. Tsukamoto, J. C. Liu, Y. P. Han, S. Govindarajan, M. M. Lai, S. Akira, and J. H. Ou.
2010. c-Jun Mediates Hepatitis C Virus Hepatocarcinogenesis through Signal Transducer and
Activator of Transcription 3 and Nitric Oxide-dependent Impairment of Oxidative DNA Repair.
Hepatology 52(2):480–92.
Mair, M., L. Blaas, C. H. Osterreicher, E. Casanova, and R. Eferl. 2011. JAK–STAT Signaling in Hepatic
Fibrosis. Front Biosci (Landmark Ed) 16:2794–811.
Meissl, K., S. Macho-Maschler, M. Müller, and B. Strobl. 2017. The Good and the Bad Faces of STAT1
in Solid Tumours. Cytokine 89:12–20.
Meraz, M. a, J. M. White, K. C. Sheehan, E. Bach, S.J. Rodig, S. Dighe, D.H. Kaplan et al. 1996. Targeted
Disruption of the Stat1 Gene in Mice Reveals Unexpected Physiologic Specificity in the JAK–STAT
Signaling Pathway. Cell 84:422–31.
Miyagi, T., M. P. Gil, X. Wang, J. Louten, W. M. Chu, and C. A. Biron. 2007. High Basal STAT4
Balanced by STAT1 Induction to Control Type 1 Interferon Effects in Natural Killer Cells. J Exp
Med 204(10):2383–96.
Moriggl, R., D.J. Topham, S. Teglund, V. Sexl, C. Mckay, D. Wang, A. Hoffmeyer et al. 1999. Stat5 Is
Required for IL-2-Induced Cell Cycle Progression of Peripheral T Cells. Immunity 10:249–59.
Mueller, K.M., J. Kornfeld, K. Friedbichler, L. Blaas, G. Egger, H. Esterbauer, P. Hasselblatt et al. 2011.
Impairment of Hepatic Growth Hormone and Glucocorticoid Receptor Signaling Causes Steatosis
and Hepatocellular Carcinoma in Mice. Hepatology 54(4):1398–409.
Murphy, K.M., and S.L. Reiner. 2002. The Lineage Decisions of Helper T Cells. Nat Rev Immunol 2
(12):933–44.
Murray, P. J. 2006. Understanding and Exploiting the Endogenous Interleukin-10/STAT3-mediated
Anti-inflammatory Response. Curr Opin Pharmacol 6(4):379–86.
Najjar, I., and R. Fagard. 2010. STAT1 and Pathogens, Not a Friendly Relationship. Biochimie 92
(5):425–44.
Naka, T., H. Tsutsui, M. Fujimoto, Y. Kawazoe, H. Kohzaki, Y. Morita, R. Nakagawa, M. Narazaki,
K. Adachi, T. Yoshimoto, K. Nakanishi, and T. Kishimoto. 2001. SOCS-1/SSI-1-Deficient NKT
Cells Participate in Severe Hepatitis through Dysregulated Cross-Talk Inhibition of IFN-γ and IL-4
Signaling in Vivo. Immunity 14(5):535–45.
Njoku, D. B., Z. Li, N. D. Washington, J. L. Mellerson, V. Monica, R. Sharma, and N. R. Rose. 2009.
Suppressive and Pro-inflammatory Roles for IL-4 in the Pathogenesis of Experimental
Drug-induced Liver Injury. Eur J Immunol 39(6):1652–63.
Numata, K., M. Kubo, H. Watanabe, K. Takagi, H. Mizuta, S. Okada, S. L. Kunkel, T. Ito, and
A. Matsukawa. 2007. Overexpression of Suppressor of Cytokine Signaling-3 in T Cells Exacerbates
Acetaminophen-induced Hepatotoxicity. J Immunol 178(6):3777–85.
Ogata, H., T. Chinen, T. Yoshida, I. Kinjyo, G. Takaesu, H. Shiraishi, M. Iida, T. Kobayashi, and
A. Yoshimura. 2006. Loss of SOCS3 in the Liver Promotes Fibrosis by Enhancing STAT3-mediated
TGF-Beta1 Production. Oncogene 25(17):2520–30.
Orlent, H., H. Reynaert, S. Bourgeois, V. Dideberg, M. Adler, I. Colle, S. De Maeght, W. Laleman,
P. Michielsen, C. Moreno, J. P. Mulkay, P. Starkel, and J. Delwaide. 2011. IL28B Polymorphism and
the Control of Hepatitis C Virus Infection: Ready for Clinical Use? Acta Gastroenterol Belg 74
(2):317–22.
156 JAK-STAT Signaling in Diseases

Park, O., H. Wang, H. Weng, L. Feigenbaum, H. Li, S. Yin, S.H. Ki et al. 2011. In Vivo Consequences of
Liver-Specific Interleukin-22 Expression in Mice: Implications for Human Liver Disease
Progression. Hepatology 54(1):252–61.
Ruff-Jamison, S., K. Chen, and S. Cohen. 1993. Induction by EGF and Interferon-gamma of Tyrosine

Phosphorylated DNA Binding Proteins in Mouse Liver Nuclei. Science 261(5129):1733–36.

Ryan, P. M., M. Bourdi, M. C. Korrapati, W. R. Proctor, R. A. Vasquez, S. B. Yee, T. D. Quinn,

M. Chakraborty, and L. R. Pohl. 2012. “Endogenous Interleukin-4 Regulates Glutathione Synth­


esis Following Acetaminophen-induced Liver Injury in Mice.” Chem Res Toxicol 25(1):83–93. doi:
10.1021/tx2003992.
Sallam, A. M., A. Esmat, and A. B. Abdel-Naim. 2018. Cucurbitacin-B Attenuates CCl4-induced
Hepatic Fibrosis in Mice through Inhibition of STAT-3. Chem Biol Drug Des 91(4):933–41.
Schindler, C. 1999. Cytokines and JAK–STAT Signaling. Exp Cell Res 253(1):7–14.
Schroder, K., P. J. Hertzog, T. Ravasi, and D. A. Hume. 2004. Interferon- Y: An Overview of Signals,
Mechanisms and Functions. J Leukoc Biol 75:163–89.
Seki, E., and D. A. Brenner. 2015. Recent Advancement of Molecular Mechanisms of Liver Fibrosis.
J Hepatobiliary Pancreat Sci 22(7):512–18.
Shrivastava, S., E. G. Meissner, E. Funk, S. Poonia, V. Shokeen, A. Thakur, B. Poonia, S. K. Sarin,
N. Trehanpati, and S. Kottilil. 2016. Elevated Hepatic Lipid And Interferon Stimulated Gene Expres­
sion in HCV GT3 Patients Relative to Non-alcoholic Steatohepatitis. Hepatol Int 10(6):937–46.
Siebler, J., S. Wirtz, S. Klein, M. Protschka, M. Blessing, P.R. Galle, and M.F. Neurath. 2003. A Key
Pathogenic Role for the STAT1/T-Bet Signaling Pathway in T-Cell-Mediated Liver Inflammation.
Hepatology 38(6):1573–80.
Simon-Holtorf, J., H. Mönig, H. J. Klomp, A. Reinecke-Lüthge, U.R. Fölsch, and S. Kloehn. 2006.
Expression and Distribution of Prolactin Receptor in Normal, Fibrotic, and Cirrhotic Human
Liver. Exp Clin Endocrinol Diabetes 114(10):584–89.
Sobrevals, L., C. Rodriguez, J. L. Romero-Trevejo, G. Gondi, Ĩ. Monreal, A. Pãneda, N. Juanarena,
S. Arcelus, N. Razquin, L. Guembe, G. González-Aseguinolaza, and J. Prieto. 2010. Insulin-like
Growth Factor I Gene Transfer to Cirrhotic Liver Induces Fibrolysis and Reduces Fibrogenesis
Leading to Cirrhosis Reversion in Rats. Hepatology 51(3):912–21.
Subleski, J. J., V. L. Hall, T. C. Back, J. R. Ortaldo, and R. H. Wiltrout. 2006. Enhanced Antitumor
Response by Divergent Modulation of Natural Killer and Natural Killer T Cells in the Liver.
Cancer Res no 66(22):11005–122. doi:10.1158/0008-5472.CAN-06-0811.
Subramaniam, S. V., R. S. Cooper, and S. E. Adunyah. 1999. Evidence for the Involvement of JAK/STAT
Pathway in the Signaling Mechanism of Interleukin-17. Biochem Biophys Res Commun 262(1):14–19.
Sun, J. 2010. Matrix Metalloproteinases and Tissue Inhibitor of Metalloproteinases Are Essential for the
Inflammatory Response in Cancer Cells. J Signal Transduct 2010:985132.
Sun, R., and B. Gao. 2004. Negative Regulation of Liver Regeneration by Innate Immunity (Natural
Killer Cells/Interferon-γ). Gastroenterology 127(5):1525–39.
Sun, R., O. Park, N. Horiguchi, S. Kulkarni, W. I. Jeong, H.Y. Sun, S. Radaeva, and B. Gao. 2006. STAT1
Contributes to DsRNA Inhibition of Liver Regeneration after Partial Hepatectomy in Mice.
Hepatology 44(4):955–66.
Tanabe, J., A. Izawa, N. Takemi, Y. Miyauchi, Y. Torii, H. Tsuchiyama, T. Suzuki, S. Sone, and K. Ando.
2007. Interferon-β Reduces the Mouse Liver Fibrosis Induced by Repeated Administration of
Concanavalin A via the Direct and Indirect Effects. Immunology 122(4):562–70.
Tang, L. Y., M. Heller, Z. Meng, L. R. Yu, Y. Tang, M. Zhou, and Y. E. Zhang. 2017. Transforming
Growth Factor-beta (TGF-beta) Directly Activates the JAK1-STAT3 Axis to Induce Hepatic
Fibrosis in Coordination with the SMAD Pathway. J Biol Chem 292(10):4302–12.
Torisu, T., M. Nakaya, S. Watanabe, M. Hashimoto, H. Yoshida, T. Chinen, R. Yoshida, F. Okamoto,
T. Hanada, K. Torisu, G. Takaesu, T. Kobayashi, H. Yasukawa, and A. Yoshimura. 2008.
Suppressor of Cytokine Signaling 1 Protects Mice against Concanavalin A-Induced Hepatitis by
Inhibiting Apoptosis. Hepatology 47(5):1644–54.
Toyonaga, T., O. Hino, S. Sugai, S. Wakasugi, K. Abe, M. Shichiri, and K. Yamamura. 1994. Chronic
Active Hepatitis in Transgenic Mice Expressing Interferon-Gamma in the Liver. Proc Natl Acad Sci
USA 91(2):614–18.
JAK-STAT Signaling in Liver Fibrosis 157

Wang, G.L., E. Salisbury, X. Shi, L. Timchenko, E.E. Medrano, and N.A. Timchenko. 2008. HDAC1
Cooperates with C/EBPα In the Inhibition of Liver Proliferation in Old Mice. J Biol Chem 283
(38):26169–788.
Wang, H., F. Lafdil, X. Kong, and B. Gao. 2011a. Signal Transducer and Activator of Transcription 3 in
Liver Diseases: A Novel Therapeutic Target. Int J Biol Sci 7(5):536–50.
Wang, H., F. Lafdil, L. Wang, O. Park, S. Yin, J. Niu, A. M. Miller, Z. Sun, and B. Gao. 2011b.
Hepatoprotective versus Oncogenic Functions of STAT3 in Liver Tumorigenesis. Am J Pathol 179
(2):714–24.
Wang, H., F. Lafdil, L. Wang, S. Yin, D. Feng, and B. Gao. 2011c. Tissue Inhibitor of Metalloproteinase
1 (TIMP-1) Deficiency Exacerbates Carbon Tetrachloride-induced Liver Injury and Fibrosis in
Mice: Involvement of Hepatocyte STAT3 in TIMP-1 Production. Cell Biosci 11:14.
Wang, K. S., J. Ritz, and D. A. Frank. 1999. IL-2 Induces STAT4 Activation in Primary NK Cells and
NK Cell Lines, but Not in T Cells. J Immunol 162(1):299–304.
Wang, W., H. Guo, H. Li, Y. Yan, C. Wu, X. Wang, X. He, and N. Zhao. 2018. Interleukin-35
Gene-Modified Mesenchymal Stem Cells Protect Concanavalin A-Induced Fulminant Hepatitis
by Decreasing the Interferon Gamma Level. Hum Gene Ther 29(2):234–41.
Wang, Y., A. Qu, and H. Wang. 2015. Signal Transducer and Activator of Transcription 4 in Liver
Diseases. Int J Biol Sci 11(4):448–55.
Wen, J., Y. Zhou, J. Wang, J. Chen, W. Yan, J. Wu, J. Yan et al. 2017. Interactions between Th1 Cells and
Tregs Affect Regulation of Hepatic Fibrosis in Biliary Atresia through the IFN- 3/STAT1 Pathway.
Cell Death Differ 24(6):997–1006.
Weng, H. L., Y. Liu, J. L. Chen, T. Huang, L. J. Xu, P. Godoy, J. H. Hu, C. Zhou, F. Stickel, A. Marx,
R. M. Bohle, V. Zimmer, F. Lammert, S. Mueller, M. Gigou, D. Samuel, P. R. Mertens,
M. V. Singer, H. K. Seitz, and S. Dooley. 2009. The Etiology of Liver Damage Imparts Cytokines
Transforming Growth Factor Β1 or Interleukin-13 as Driving Forces in Fibrogenesis. Hepatology
50(1):230–43.
Weng, S. Y., X. Wang, S. Vijayan, Y. Tang, Y. O. Kim, K. Padberg, T. Regen, O. Molokanova, T. Chen,
T. Bopp, H. Schild H. F. Brombacher, J. R. Crosby, M. L. McCaleb, A. Waisman, E. Bockamp, and
D. Schuppan. 2018. IL-4 Receptor Alpha Signaling through Macrophages Differentially Regulates
Liver Fibrosis Progression and Reversal. EBioMedicine 29:92–103.
Woelfle, J., J. Billiard, and P. Rotwein. 2003. Acute Control of Insulin-like Growth Factor-I Gene
Transcription by Growth Hormone through Stat5b. J Biol Chem 278(25):22696–702.
Wurster, A. L., T. Tanaka, and M. J. Grusby. 2000. The Biology of Stat4 and Stat6. Oncogene 19
(21):2577–84.
Xu, H., Z. Fan, W. Tian, and Y. Xu. 2016. Protein Inhibitor of Activated STAT 4 (PIAS4) Regulates Liver
Fibrosis through Modulating SMAD3 Activity. J Biomed Res 30(6):496–501.
Yang, Y., X. Q. Wu, W. X. Li, H. M. Huang, H. D. Li, X. Y. Pan, X. F. Li, C. Huang, X. M. Meng,
L. Zhang, X. W. Lv, H. Wang, and J. Li. 2018. PSTPIP2 Connects DNA Methylation to
Macrophage Polarization in CCL4-induced Mouse Model of Hepatic Fibrosis. Oncogene 37
(47):6119–35.
Younossi, Z. M., R. Loomba, Q. M. Anstee, M. E. Rinella, E. Bugianesi, G. Marchesini,
B. A. Neuschwander-Tetri, L. Serfaty, F. Negro, S. H. Caldwell, V. Ratziu, K. E. Corey,
S. L. Friedman, M. F. Abdelmalek, S. A. Harrison, A. J. Sanyal, J. E. Lavine, P. Mathurin,
M. R. Charlton, Z. D. Goodman, N. P. Chalasani, K. V. Kowdley, J. George, and K. Lindor. 2018.
Diagnostic Modalities for Nonalcoholic Fatty Liver Disease, Nonalcoholic Steatohepatitis, and
Associated Fibrosis. Hepatology 68(1):349–60.
Yu, J. H., B. Zhu, M. Wickre, G. Riedlinger, W. Chen, A. Hosui, W. R. Gertraud, and
L. Hennighausen. 2011. The Transcription Factors STAT5A and STAT5B Negatively Regulate
Cell Proliferation through the Activation of Cdkn2b and Cdkn1a Expression. Hepatology 52
(5):1808–18.
Zhang, H., F. Chen, X. Fan, C. Lin, Y. Hao, H. Wei, W. Lin, Y. Jiang, and F. He. 2017. Quantitative
Proteomic Analysis on Activated Hepatic Stellate Cells Reversion Reveal STAT1 as a Key Reg­
ulator between Liver Fibrosis and Recovery. Sci Rep 7:1–10.
158 JAK-STAT Signaling in Diseases

Zhang, L., N. Jilg, R. Shao, W. Lin, and D. Fusco. 2011. IL28B Inhibits Hepatitis C Virus Replication
through the JAK–STAT Pathway. J Hepatol 55(2):289–98.
Zhu, R., S. Diem, L. M. Araujo, A. Aumeunier, J. Denizeau, E. Philadelphe, D. Damotte, M. Samson,
P. Gourdy, M. Dy, E. Schneider, and A. Herbelin. 2007. The Pro-Th1 Cytokine IL-12 Enhances
IL-4 Production by Invariant NKT Cells: Relevance for T Cell-mediated Hepatitis. J Immunol 178
(9):5435–42.
10
Renal Disorders: Involvement of JAK-STAT
Pathway

Yuji Nozaki
Department of Hematology and Rheumatology
Kindai University School of Medicine
Osaka-Sayama, Japan

10.1 Introduction
Many signaling pathways have been shown to be involved in the progression of renal disease in
humans and in animal models, usually due to persistent activation of these pathways. Perhaps the
best well known of such pathways include certain protein kinase pathways, such as protein kinase
C and mitogen-activated kinase pathways, which lead to cellular growth and fibrosis, and those,
such as angiotensin (ANG) II and transforming growth factor-β (TGF-β)/SMAD signaling path­
ways, which enhance fibrosis leading to kidney scarring (Chen, Chen, and Harris 2012; Lan and
Chung 2012; Li and Gobe 2006; Siragy and Carey 2010). In addition, mesangial cells possess both
contractile and mitogenic properties and contribute to the physiological regulation of glomerular
dynamics (Johnson et al. 1993; Wenzel et al. 1995). Growth factors have a maladaptive role in the
glomerular damage accompanying experimental and human glomerulonephritis.
One of the major pathways that responds to and transduces inflammatory signals is the Janus
kinase–signal transducer and activator of transcription (JAK-STAT) pathway. The JAK-STAT
pathway transmits signals from extracellular ligands, including many cytokines and chemokines
as well as growth factors and hormones, directly to the nucleus to induce a variety of cellular
responses (Lu et al. 2009). While many of these responses are present and best characterized in
lymphoid cells, they have also been reported in many other cell types including kidney parench­
ymal cells such as podocytes, mesangial cells, and tubular cells (Berthier et al. 2009; Choudhury,
Ghosh-Choudhury, and Abboud 1998). Given the role of JAK-STAT activation in response to
cytokines and chemokines and the unequivocal role of inflammation in promoting progressive
kidney injury, it is not surprising that JAK-STAT activation is involved in the pathogenesis of
renal disease as well as of acute kidney injury. Since its discovery, JAK-STAT pathway has been
intensively studied. Because of its strong immunomodulatory function, JAK-STAT pathway and
its components are good candidates for immunological intervention for disease control. Indeed,
several clinical trials administering JAK inhibitor have been performed for various diseases such
as chronic kidney disease (CKD), rheumatoid arthritis (RA), psoriasis, inflammatory bowel
disease, and atopic dermatitis (Boyle et al. 2015; Brosius, Tuttle, and Kretzler 2016; Levy et al.
2015; Olivera, Danese, and Peyrin-Biroulet 2017).
There is considerable interest in JAK-STAT as a therapeutic target, especially as a treatment for
renal disease. CKD in renal disorders is one of the leading causes of morbidity and mortality in
patients with diabetic nephropathy, lupus nephritis, and kidney transplantation, and ANG II and
inflammatory cytokines have been implicated in the pathogenesis of maladaptive growth and
inflammation in renal tissues in these patients (Ling, Seal, and Eaton 1993; Wang et al. 2010).

159
160 JAK-STAT Signaling in Diseases

In this review, the JAK-STAT signaling mechanism and how the inhibition of JAK-STAT expres­
sion aids in the development of effective treatments for the renal diseases have been discussed.

10.2 JAK-STAT Signaling


The JAK-STAT pathway is the major signaling cascade downstream of type I and type II
cytokine receptors. JAKs have been shown to be an effective therapeutic target for various
autoimmune and inflammatory diseases driven by cytokines (Clark, Flanagan, and Telliez 2014;
O’Shea et al. 2015). JAKs are cytoplasmic tyrosine kinases that participate in the signaling of
a broad range of cell surface receptors, particularly members of the cytokine receptor common
gamma (cg) chain family (Vincenti and Kirk 2008). There are four mammalian JAKs: JAK1, 2, 3,
and tyrosine kinase 2 (Tyk2). The activation of JAK occurs by a ligand-receptor interaction,
which results in signaling through the phosphorylation of cytokine receptors and the creation of
docking sites for signaling proteins known as STAT (Podder and Kahan 2004). JAKs catalyze
STAT phosphorylation, which facilitates STAT dimerization and nuclear transport. The result is
the regulation of gene expression and transcription (Podder and Kahan 2004; Wiik et al. 1994).
Although JAK proteins are structurally related, their differences in activation and downstream
effects allow for a high degree of specificity. JAK1 is activated by the ligands binding to the class II
receptors (interferon (IFN)-α/β, IFN-γ, and interleukin (IL)-10) and the γc receptors (IL-2, IL-4,
IL-7, IL-9, IL-15, and IL-21) (Figure 10.1). JAK2 is activated mostly by thrombopoietin receptor,
IL-3, granulocyte-macrophage-colony-stimulating factor (GM-CSF), and IFN-γ. While the majority
of JAKs are ubiquitously expressed, JAK3 expression appears to be generally restricted to
hematopoietic lineages and vascular muscle cells and JAK3 plays a critical role in lymphocyte
development and function. Tyk2 mediates mostly the signaling induced by IL-12. JAK3, unlike
the ubiquitous expression of other JAK subtypes, has a restricted tissue distribution and is
primarily found on hematopoietic cells and uniquely associates with the cg chain (Witthuhn
et al. 1994). The importance of this signaling pathway is evident when one considers that mice
and humans, with genetic absence or mutation in either the cg subunit or JAK3, express defects
in lymphoid development that give rise to a severe combined immunodeficiency syndrome
phenotype (Buckley 2004).
Since the JAK-STAT pathways play major activating roles in a variety of disease processes,
there has been a robust effort to develop specific inhibitors of this pathway. As it is relatively easy
to identify inhibitors of protein kinases, development of JAK inhibitors has received most
attention. At present, three JAK inhibitors have received FDA approval for clinical use. Ruxoli­
tinib (Jakafi, Incyte) is a potent inhibitor of both JAK1 and JAK2, and it received FDA approval
in November 2011 for the treatment of polycythemia vera and myelofibrosis (Verstovsek et al.
2010). Tofacitinib (Xeljanz, Pfizer) was initially designed to be a specific inhibitor of JAK1 and
JAK3 kinase and it received FDA approval in November 2012, and therefore has been used as an
immunosuppressant in transplantation and for the treatment of autoimmune diseases (Borie et al.
2003). Baricitinib as a JAK1 and JAK2 inhibitor (Olumiant, Eli Lilly) received FDA approval for
the treatment of moderately to severely active RA in June 2018. Several other JAK inhibitors have
been developed as immunosuppressive agents for RA and other autoimmune diseases. For
example, upadacitinib (a JAK1 inhibitor), filgotininib (a JAK1 inhibitor), and peficinib (a JAK1
inhibitor) have been shown to be effective in the treatment of arthritis (Burmester et al. 2018;
Westhovens et al. 2017). Therefore, it is likely that additional JAK inhibitors will become clinically
available in the next few years.
STAT genes are also essential for proper immune function, and loss-of-function mutations in
these proteins can be associated with immunodeficiency syndromes (Holland, DeLeo, and Elloumi
2007). Certain JAK-STAT polymorphisms are associated with an increased risk of developing
autoimmune diseases (Schwartz et al. 2016). These observations have also led to the development
of JAK inhibitors for the treatment of human disease (Clark, Flanagan, and Telliez 2014).
Renal Disorders

FIGURE 10.1 JAK inhibitions and immune regulation by JAK pathway. Inhibition of JAK3 would affect signaling mediated by only the common gamma chain-associated cytokine
receptors (IL-2R, IL-4R, IL-9R, and IL-21R) and regulate T cell, NK cell, and B cell function. On the other hand, JAK2 or TYK2 inhibition would influence multiple cytokine
receptor signaling pathways.
Abbreviation: EPO, erythropoietin: GH, growth hormone: TPO, thrombopoietin: GM-CSF, granulocyte macrophage colony-stimulating factor: PRL, prolactin receptor: CNTFR,
ciliary neutrophic factor receptor: IL, interleukin: IFN, interferon: NK, natural killer.
161
162 JAK-STAT Signaling in Diseases

10.3 JAK Inhibitor


10.3.1 Tofacitinib
Tofacitinib is a first-generation JAK inhibitor that inhibits JAK1 and JAK3 (with greater potency
than JAK2) (Changelian et al. 2003; Furumoto and Gadina 2013). Since common γ chain-using
cytokines use JAK1 and JAK3, tofacitinib efficiently blocks their signaling cascades. Tofacitinib
also blocks gp130-using cytokines as well as signaling downstream of type I/II IFNs. Tofacitinib
may, therefore, have diverse effects on cytokines and cells that contribute to SLE pathogenesis,
including subsets of CD4+ T cells (helper T cells including Th1 and pathogenic Th17 cells), CD8+
T cells, B cells, and innate immune cells (Ghoreschi, Jesson, and Li 2011; Okiyama et al. 2014;
Xing et al. 2014). Tofacitinib has been reported to increase HDL and a previous report indicated
that this drug can restore lipoprotein homeostasis and improve cholesterol efflux capacity in RA
patients (Boers et al. 2003; Charles-Schoeman et al. 2015). The effects of tofacitinib on
cholesterol transport were proposed to be resulting from an increase in cholesterol ester
production rate through augmented activity of the lecithin-cholesterol acyltransferase (Charles-
Schoeman et al. 2015). Given the changes observed in lipoprotein profiles of tofacitinib-treated
mice and, in particular, the reduction in the fraction of free cholesterol—which has recently
been shown to be pro-atherogenic—the lipoprotein changes induced by tofacitinib could be
an additional mechanism by which JAK inhibition could be vasculoprotective in systemic
autoimmunity (Thacker et al. 2015).

10.3.2 Baricitinib
Baricitinib is a novel oral molecule that potently inhibits JAK1 and JAK2 (with greater potency
than JAK3 and TyK2) (Richez et al. 2017). The drug was developed as an immunosuppressive
agent for the treatment of RA, atopic dermatitis, and systemic lupus. Its development in psoriasis
and diabetic chronic disease has been halted. In vitro, baricitinib inhibited generation of type I and
type II IFNs (IFNα and IFNγ), the common γ-chain cytokines (IL-15 and IL-21), and IL-6 and IL­
27. Baricitinib also showed some decrease in potency for IL-10, IL-12, IL-23, and erythropoietin
(EPO). Baricitinib demonstrated dose- and time-dependent inhibition of cytokine-induced pSTAT3,
with maximal inhibition of pSTAT3 occurring 1–2 h after administration for all doses (Shi et al.
2014). Efficacy in rat adjuvant-induced arthritis correlated with the level of pSTAT3 in a dose- and
time-dependent manner. Efficacy was observed as soon as pSTAT3 levels were inhibited by 15% and
even more than 24 h later, suggesting that baricitinib possesses significant activity even in the
absence of complete and continuous pathway inhibition (Fridman et al. 2010). In cell-based assays
using human T cells, the drug potently inhibited phosphorylation of STAT3 and subsequent
production of monocyte chemoattractant protein-1, as well as IL-23, and induced STAT3 phos­
phorylation and the subsequent production of IL-17 and IL-22 (Markham 2017). Furthermore,
JAK1 and JAK2 inhibition, and not JAK3 alone, has been shown to control STAT3-dependent
signaling pathway (Migita et al. 2013).

10.3.3 Ruxolitinib
Ruxolitinib is unique among JAK kinase inhibitors in clinical development as it potently inhibits
JAK1 in addition to JAK2 for the treatment of polycythemia vera and myelofibrosis (Verstovsek et al.
2010). JAK1 is hyperactivated in the peripheral blood of myelofibrosis patients (Quintás-Cardama
et al. 2010). Furthermore, Jedidi et al. demonstrated that cytokines capable of signaling through
JAK1 can convey resistance to JAK2-targeted therapies (Jedidi et al. 2009). JAK2 also mediates
cytokine signaling from EPO, thrombopoietin (TPO), and GM-CSF receptors (Witthuhn et al. 1993).
Germline deletion of JAK2 causes embryonic lethality by deficient definitive erythropoiesis, which
was confirmed by a conditional JAK2 allele and germline Cre-recombinase (Krempler et al. 2004;
Renal Disorders 163

Neubauer et al. 1998; Parganas et al. 1998). Conditional JAK2 loss in hematopoietic stem cells/
progenitor cells induces anemia and thrombocytopenia, whereas deletion of the TPO receptor MPL
and MPL in platelets leads to thrombocytosis (Grisouard et al. 2014; Lannutti et al. 2009; Ng et al.
2014; Park et al. 2013; Pikman et al. 2006; Tiedt et al. 2009). Somatic mutations in the JAK2 pathway,
including JAK2V617F, JAK2 exon 12 mutations, and in the MPL, are observed in myeloproliferative
neoplasms (MPN) patients. JAK kinase inhibitors attenuate splenomegaly and constitutional symp­
toms in MPN but cannot induce molecular responses (Baxter et al. 2005; Harrison et al. 2012; James
et al. 2005; Kralovics et al. 2005; Levine et al. 2005; Scott et al. 2007; Verstovsek et al. 2012). JAK
inhibitor therapy results in the variable development of anemia and thrombocytopenia preventing
dose intensification and maximal target inhibition (Mesa and Cortes 2013). Cytopenias have been
considered on-target effects of JAK2 inhibitors inseparable from therapeutic efficacy. However, the
risk of thrombocytopenia varies among JAK2 inhibitors, suggesting thrombocytopenia may not be
a direct consequence of JAK2 inhibition (Tam and Verstovsek 2013). Thus, there may be both cell
autonomous and non-autonomous benefits of inhibiting JAK1 as well as JAK2.

10.3.4 Next Generation of JAK Inhibitors


The various clinical trials in the new JAK inhibitors such as filgotinib (Galapagos/Gilead),
upadacitinib (AbbVie), peficitinib (Smyraf, Astellas), and decernotinib (Vertex) are running in
RA (Burmester et al. 2018; Genovese et al. 2016, 2017; Kavanaugh et al. 2017). These data are
seemed to be very promising for treating RA. Filgotinib demonstrated a selective inhibition of
JAK1 and JAK2 over JAK3 and Tyk2, and specifically in whole blood assays, a selectivity of 30­
fold for JAK1 over JAK2 was revealed. Details of the in vitro assays used for determining JAK
selectivity are discussed by Van Rompaey et al. in their preclinical work on filgotinib (Van et al.
2013). Upadacitinib demonstrated 74- and 58-fold greater selectivity for JAK1 over JAK2 and
JAK3, respectively (Mohamed et al. 2016). Peficitinib has moderate selectivity for JAK3 and
inhibited JAK1 and JAK3 with 50% inhibitory concentrations of 3.9 and 0.7 nM, respectively; it
has also shown 7.1-fold selectivity for JAK3 relative to JAK2. Peficitinib phase III has performed
in south-east Asia only. Decernotinib, in first evaluation, was demonstrated to be a selective and
potent inhibitor of JAK3 in vitro and modulated pro-inflammatory responses in models of
immune-mediated diseases, such as collagen-induced arthritis and delayed-type hypersensitivity
(Mahajan et al. 2015). Both JAK1 and JAK3 inhibition apparently are effective and one needs to
explore the mechanistic pathways leading to efficacy. Just labeling for certain specificity is not
appropriate when one looks at the differences in ex vivo results and in vitro cellular pharmacology
from baricitinib to upadacitinib, filgotinib, and tofacitinib (McInnes, Higgs, and Lee 2017). JAK
inhibitors display different in vitro pharmacological profiles, which, coupled to their in vivo
pharmacokinetics, suggests that they modulate distinct cytokine pathways to differing degrees
and durations over 24 h. Baricitinib and filgotinib inhibited JAK1/3 signaling to a lesser extent
than upadacitinib and tofacitinib. Differences between the drugs are certainly there and the
differential clinical relevance needs specific study. In renal disorders, further studies also need to
examine whether new JAK inhibitors could be effective on the progression of renal damage.

10.4 Renal Disease


10.4.1 Diabetic Nephropathy
Diabetic nephropathy is the most common cause of end-stage renal disease (ESRD) in many
other countries today (Saran et al. 2015). Furthermore, its incidence continues to increase
because of the epidemic of type 2 diabetes, despite the improvement in the rates of progressive
diabetic nephropathy (Saran et al. 2015). Diabetic nephropathy is characterized by sustained
inflammation that promotes and directs much of the chronic injury process (Wada and Makino
164 JAK-STAT Signaling in Diseases

2013). One of the major pathways that transduce inflammatory signals in diabetic nephropathy
is the JAK-STAT pathway. The JAK-STAT pathway transmits signals from extracellular
ligands, including many cytokines and chemokines as well as growth factors and hormones,
directly to the nucleus to induce a variety of cellular responses (O’Shea and Plenge 2012). While
many of these responses are best characterized in lymphoid cells, they have also been reported
in intrinsic kidney cells such as podocytes, mesangial cells, and tubular cells (Berthier et al.
2009; Choudhury, Ghosh-Choudhury, and Abboud 1998). Gene and protein expression studies
of kidney biopsies from people with early- and late-stage diabetic nephropathy have shown
increased activation and expression of the JAK-STAT signaling pathway across the spectrum of
diabetic nephropathy (Berthier et al. 2009; Woroniecka et al. 2011). In particular, increased
expression and activity of JAK1 and JAK2 appear to promote diabetic nephropathy develop­
ment and progression (Brosius and He 2015; Navarro-González et al. 2011; Zhang et al. 2017).
Moreover, studies have suggested interactions between JAK-STAT and angiotensin signaling,
including evidence for activating JAK2 (Marrero et al. 2006). Marrero MB et al. also reported
that activation of the JAK-STAT pathway is essential for ANG II-induced growth of glomer­
ular mesangial cells (Amiri et al. 2002; Wang et al. 2002). In addition, they have also shown
that high glucose augments ANG II-induced activation of the JAK-STAT pathway in rat kidney
glomeruli and that this pathway contributes importantly to production of the cytokine TGF-β
as well as the extracellular matrix proteins—collagen IV and fibronectin (Banes et al. 2004;
Wang et al. 2002). Furthermore, Ha H et al. have also demonstrated that high glucose induces
intracellular production of reactive oxygen species (ROS), which have been shown to augment
ANG II-induced signaling cascades and activate JAK and STAT proteins (Ha and Lee 2000;
Simon et al. 1998; Ushio-Fukai et al. 1999). Figure 10.2 show the schematic depiction of high
glucose-induced augmentation of ANG II-induced growth via the JAK-STAT pathway in
glomerular mesangial cells. In the context of diabetic complications, evaluating the molecular
mechanisms responsible for modulating ANG II-induced activation of the JAK-STAT pathway
in mesangial cells by high glucose is an important area. The mechanisms by which high glucose
promotes JAK2 activation may be related to activation of JAK2 by ROS, and ROS are induced

FIGURE 10.2 Schematic depiction of high glucose-induced augmentation of ANG II-induced growth via the

JAK-STAT pathway in glomerular mesangial cells.

Abbreviations. ROS, reactive oxygen species: GMC, glomerular mesangial cell: ANG II, angiotensin II: STAT, signal

transducer and activator of transcription: P, phosphorylation: JAK2, Janus kinase2: TGF-β, transforming growth factor-β.

Renal Disorders 165

by high glucose in mesangial cells (Ha and Lee 2000). A previous report has shown that high glucose,
via the polyol pathway, induces a rapid increase in intracellular ROS, such as H2O2, which stimulates
intracellular signaling events similar to those activated by ANG II, including phosphorylation of
growth-promoting kinases such as JAK2 (Shaw et al. 2003). The polyol pathway generates ROS (Ha
and Lee 2000; Shaw et al. 2003), which can then act as signaling mediators in the activation of
downstream mitogenic pathways, such as the JAK-STAT cascade (Simon et al. 1998). It has also
recently been reported that ANG II induces a rapid increase in intracellular H2O2, via NAD(P)H
oxidase, which subsequently activates growth-related responses (Ushio-Fukai et al. 1999). In the
addition, Bhat GJ et al. has shown that high glucose altered activation of the JAK-STAT pathway
in vivo through ANG II in rat kidney glomeruli by inducing phosphorylation of JAK2 kinase and
STAT proteins, namely, STAT1, STAT3, and STAT5A/B (Bhat et al. 1994, 1995; Chappey et al.
1997). The authors also reported that the streptozotocin-induced diabetic rats treated with the JAK2
inhibitor tyrphostin showed a significant reduction in urine output as well as a significant reduction in
fluid intake. Furthermore, treatment with tyrphostin lowered systolic blood pressure and significantly
reduced urinary protein excretion. These results provide further support for the hypothesis that the
JAK2 contributes importantly to both acute and chronic glomerular dysfunction in diabetic nephro­
pathy. Berthier CC et al. also found that expression of JAK2 protein in proximal tubule cells was high
in patients with progressive kidney disease compared with that in patients without kidney disease and
even compared with that in a few individuals with other glomerular diseases (Berthier et al. 2009). Li
X et al. reported that Paeoniflorin attenuates renal lesions in diabetic mice and the protective effects
may be associated with the prevention of macrophage infiltration and inhibition of the JAK2–STAT3
signaling pathway (Li et al. 2018). Other important regulators of JAK-STAT signaling are the
suppressors of cytokine signaling (SOCS). SOCS1 and SOCS3 are downstream transcriptional targets
of the JAK-STAT signaling pathway that bind to JAK proteins and interfere with their activity in
a negative-feedback manner (Croker, Kiu, and Nicholson 2008). Rats injected with recombinant
SOCS1 and SOCS3 adenovirus showed evidence of reduced JAK-STAT activation and amelioration
in early diabetic changes (Õrtiz-Munoz et al. 2010). A previous study also reported that STAT3
acetylation is increased in both mouse and human diabetic kidneys (Liu et al. 2014). This has
functional implications because acetylation of STAT proteins enhances STAT dimerization, which is
critical for the translocation of STATs to the nucleus and their ability to modulate gene transcription
(Zhuang 2013). Treatment with inhibitors that block acetylation-mediated association of STAT3
reduced both proteinuria and kidney injury in the db/db mouse model of diabetes.
Baricitinib, known as Olumiant, is a novel oral molecule that potently inhibits JAK1 and JAK2
(with greater potency than JAK3 and TyK2) (Richez et al. 2017). In diabetic nephropathy, baricitinib
treatment resulted in a reduction in albuminuria at 3 and 6 months (Tuttle et al. 2018). After 4 weeks
of study drug wash-out, the albuminuria reduction was sustained in the medium- and high-dose
baricitinib groups. The reduction in albuminuria was approximately 40% versus placebo in the
highest-dose baricitinib group. Importantly, there was evidence of target engagement and
a reduction in renal inflammation with baricitinib treatment, as assessed by urinary levels of IFN-γ­
induced protein 10 and plasma levels of TNF receptor 2 (TNFR2). Elevated plasma TNFR2 and
TNFR1 levels have been reported as the biomarkers with high prognostic power for progression of
diabetic nephropathy in both type 1 and type 2 diabetic populations (Gohda et al. 2012; Niewczas
et al. 2012). However, there was a statistically significant decrease in hemoglobin compared to placebo
in the high-dose baricitinib group at 6 months because EPO signaling is dependent on JAK2
activation (Kuhrt and Wojchowski 2015). Further studies are needed to examine the effects of
baricitinib on progression of diabetic nephropathy for the treatment as determined by harder
endpoints, such as estimated glomerular filtration (eGFR) decline, ESRD, and mortality events.

10.4.2 Lupus Nephritis


Systemic lupus erythematosus (SLE) is a chronic systemic autoimmune disorder characterized by
loss of tolerance against nuclear autoantigens, anti-double-stranded DNA antibody production,
166 JAK-STAT Signaling in Diseases

immune complex (IC) deposition, and leukocyte infiltration in many target organs, as well as
activated B and T cells (Bagavant and Fu 2009; Tsokos 2011). Lupus nephritis is an IC
glomerulonephritis categorized as one of the most serious complications of SLE and is also one
of the strongest predictors of a poor prognosis. Lupus nephritis can lead to severe proteinuria,
hypertension, chronic renal failure, and, finally, ESRD. The lupus mouse model MRL/MpJ-Faslpr
/lpr/J (MRL/lpr) recapitulates several of the clinical manifestations and immune dysregulation
observed in human SLE. Indeed, these mice develop IC glomerulonephritis, inflammatory skin
disease, aberrant T cell responses, dysregulated inflammatory cytokine synthesis, and an enhanced
type I IFN signature (Cohen and Eisenberg 1991; Perry et al. 2011). Furthermore, MRL/lpr mice
also represent a good model to study vascular dysfunction and neutrophil dysregulation char­
acteristic of lupus (Andrews, Eisenberg, and Theofilopoulos 1978). Furumoto et al. reported the
inhibition of the JAK-STAT pathway significantly ameliorated the lupus clinical phenotype and
modulated features of dysregulated innate and adaptive immunity characteristic of this disease
(Furumoto et al. 2017). Importantly, tofacitinib significantly improved vascular parameters
suggesting a potential modulatory role in cardiovascular risk in this disease. While the treatments
for SLE have improved over the last several decades, current therapies are still suboptimal,
promote significant side effects and, to this date, have not shown to consistently modify both
lupus disease activity, and the enhanced cardiovascular risk characteristic of this disease. Fur­
umoto et al. also described tofacitinib has pleiotropic beneficial effects in murine lupus, which
includes: (a) significant amelioration of clinical phenotype and organ damage; (b) modulation of
innate and adaptive immune dysregulation; and (c) improvement in vascular functions and
lipoprotein profiles. The observed effects of tofacitinib on innate and adaptive immune dysregula­
tion likely play key roles in modifying clinical responses and vasculopathy. Indeed, pharmacologic
JAK inhibition resulted in significant abrogation in IFN responses and other pro-inflammatory
cytokines considered crucial in lupus pathogenesis, neutrophil extracellular traps (NETs), and
T lymphocyte subsets associated with lupus pathogenicity. The main effects induced by tofacitinib
on adaptive immune cells were a reduction of CD8+ and dendritic (DN) T cells. This is consistent
with other inflammatory models driven by cytotoxic CD8+ lymphocytes, where JAK inhibition
strongly suppressed these responses (Craiglow and King 2015; Okiyama et al. 2014; Xing et al.
2014). In SLE, higher numbers of cytotoxic, effector CD8+ T cells are associated with disease
activity and may lead to autoantigen generation through perforin/granzyme-related pathways
(Blanco et al. 2005). DN T cells are expanded in the peripheral blood and tissues from SLE
patients as well as murine models. These cells synthesize pro-inflammatory cytokines and induce
antibody production through promotion of B-cell differentiation (Boggio et al. 2004; Crispin et al.
2008; Edgerton et al. 2009; Wong et al. 2008). Given the significant reductions in splenic DN
T cells induced by tofacitinib and the decreases in circulating cytokine levels, it is likely that some
of the beneficial effects of this drug in the lupus phenotype, including vasculopathy, are due to
repression of such responses (Usui et al. 2012). Synthesis of type I IFNs driven by endogenous
nucleic acids is considered an early event that primes the immune system for immune dysregula­
tion and development of autoimmunity. As such, strategies that target the type I IFN pathway are
actively being investigated (Kirou and Gkrouzman 2013). As type I IFNs signal through the JAK­
STAT pathway, tofacitinib blocked the biological responses to these cytokines that was reflected
in significant decreases in the type I IFN signature in these mice. It is likely that inhibition of this
pathway by tofacitinib led to pleiotropic effects limiting dysregulation of innate and adaptive
immune responses including roles in priming neutrophils to undergo NET formation, alterations
in B cell ontogeny, as well as improvements in lupus vasculopathy. NETs have been proposed to
represent an important source of immunostimulatory molecules and modified autoantigens that
can trigger loss of tolerance in predisposed hosts. Furthermore, NETs are involved in the
induction of type I IFN responses through activation of plasmacytoid DN cells. Cytokines,
autoantibodies, and immune complexes can trigger enhanced NETs formation and may be
involved in neutrophil dysregulation in SLE (Kaplan 2011). As tofacitinib treatment led to
decreases in all of these lupus-associated features, it is possible that this is an important
mechanism by which this drug improves the clinical phenotype as well as vascular dysfunction.
Renal Disorders 167

While the mechanisms leading to amelioration of the vasculopathy with tofacitinib are yet
unclear, the significant modulation that this drug promoted in both type I IFN responses and
NET formation, both considered important factors in premature endothelial damage, make it
likely that these are the key downstream pathways by which this drug was effective (Denny et al.
2010; Knight et al. 2013; Thacker et al. 2010, 2012).
The specific blockade of JAK2 could also help treat several of the pathological manifestations of
SLE, including arthritis and dermatitis, and may even help prevent or treat some comorbidities,
such as atherosclerosis, as JAK2 has demonstrated a role in the development of atherosclerotic
plaques and is necessary for EPO function in megakaryocytes (Ghoreschi, Laurence, and O’Shea
2009; López-Pedrera et al. 2010; Mazière, Conte, and Mazière 2001). Multiple cytokines have been
implicated in playing key roles during the initiation, progression, and development of SLE
including, but not limited to, IL-6, IL-12, and type I IFN (α/β) (Aringer and Smolen 2004; Chun
et al. 2007; Niewold et al. 2007; Tucci et al. 2008). All three cytokines signal via receptors controlled
by JAK kinases. Signaling of IL-6 via the IL-6R on activated B cells induces the dimerization with
gp130 and activation of the receptor-associated JAK tyrosine kinases, JAK1 and JAK2. Most
important is the role of IL-6 as this cytokine has been implicated in multiple autoimmune diseases
and directly contributes to plasma cell survival in bone marrow niches (Winter et al. 2010). In
addition, multiple studies in mouse models of SLE have repeatedly demonstrated the importance of
IL-6 in driving SLE disease manifestations (Aringer and Smolen 2004; Cash et al. 2010; Chun et al.
2007; Jeon et al. 2010; Ripley et al. 2005). The activation of JAKs triggers the phosphorylation of
IL-6R and gp130, followed by the activation of various secondary messengers and transcription
factors including STAT3, MAPKs, and Akt providing growth and proliferation signals (Silver and
Hunter 2010). Genetic ablation or polymorphisms in key suppressors of JAK-STAT signaling such
as SOCS have been implicated in increased serum IL-6 levels and risk of developing SLE in humans
(Fujimoto et al. 2004; Yu et al. 2003). In addition, JAKs play a critical role in transducing signals
from the IL-6R, and IL-6 is involved in both SLE and the maintenance of the pool of potentially
autoreactive plasma cells. Therefore, blockade of JAK signaling using a selective and potent JAK2
inhibitor could weaken the supportive effects of IL-6 on sustaining autoreactive plasma cells in
SLE. Targeting cytokine/growth factor pathways important for plasma cell generation and SLE
development is supported by literature, and targeting the IL-6 pathway and receptor for SLE
treatment is currently being tested (Febbraio, Rose-John, and Pedersen 2010; Illei et al. 2010; Silver
and Hunter 2010). Targeting of the BAFF pathway has been successful, and studies in the field are
under way to look at the role of a proliferation-inducing ligand in plasma cell generation (Town­
send, Monroe, and Chan 2010). A possible next step is to target the IL-6R pathway via the
blockade of JAK2 to better target the development and survival of autoreactive, pathogenic plasma
cells during early SLE.
However, it would be the possibility to progress the reduction of red blood cells, presumably
due to JAK2-mediated EPO inhibition; in future studies, it will be important to assess the effect
of JAK inhibitors on predisposition to infections and overall immune surveillance (Manshouri
et al. 2008).

10.4.3 Kidney Transplantation


ESRD continues to be an expanding problem facing the healthcare system. The two primary
treatment options for ESRD are dialysis and kidney transplantation. For many patients with
ESRD, kidney transplantation improves the quality of life and patient survival compared to
dialysis therapies (Wolfe et al. 1999). With the need for kidney transplant outpacing the supply of
organs, it becomes paramount that we achieve the best outcome of the renal allografts, which
includes maximizing renal allograft survival and minimizing the need for a second transplant.
However, despite improved short-term success in allograft survival that has come with improved
anti-rejection therapies, there has not been a commensurate improvement in long-term renal
allograft survival (Lodhi, Lamb, and Meier-Kriesche 2011; Meier-Kriesche et al. 2004).
168 JAK-STAT Signaling in Diseases

Chronic allograft nephropathy (CAN), a nonspecific term describing interstitial fibrosis and tubular
atrophy, remains the most common cause of late allograft loss, while cardiovascular disease continues
to be the leading cause of death with a functioning allograft post-transplantation (Ducloux et al.
2002; Jevnikar and Mannon 2008). The current challenge facing the transplant community is to
develop immunosuppression that maintains short-term success but prolongs allograft survival. The
advent of the calcineurin inhibitors (CNIs), cyclosporine (CsA) and tacrolimus, in the 1980s and
1990s, respectively, in conjunction with enhanced anti-proliferative agents such as mycophenolate
mofetil (MMF) and the current widespread use of induction agents have contributed to lower acute
rejection rates and improved 1-year kidney transplant allograft survival. The core of most current
immunosuppressive protocols is CNI therapy. CNIs also contribute to worsening hypertension,
diabetes, and dyslipidemia (Gaston et al. 2010; Mathis et al. 2004; Vincenti et al. 2007) and thus
impart negative effects on the cardiometabolic risk profile. CNI minimization strategies to avoid
nephrotoxicity may also have contributed to the development of chronic antibody-mediated rejection,
which is an important cause of late graft failure (Karaman et al. 2008).
A previous study showed that tofacitinib has high affinity for JAK3, with little binding to
unrelated kinases (Changelian et al. 2003). The in vivo effect of tofacitinib was first investigated in
animal models of organ graft rejection. Efficacy could be shown for preventing heart or kidney
rejection after transplantation without observing metabolic abnormalities or severe side effects due
to immunosuppression (Kudlacz et al. 2004). Marked reduction in lymphocyte subsets was observed
in a dose-dependent manner but was transient since after dosing cessation lymphocyte numbers
began to normalize (Conklyn et al. 2004). In cynomolgus monkeys, oral dosing of JAK3 mainly
reduced numbers of natural killer (NK) cells and effector memory CD8+ T cells (Borie et al. 2005).
Prevention of graft rejection and prolongation of kidney allograft survival in cynomolgus monkeys
has been suggested to be related to the decrease of NK cells and T cells and lower production levels
of IFN-γ into facitinib-treated animals (Paniagua et al. 2005; Rousvoal et al. 2006). JAK3 inhibition
with this compound was also effective in a transplantation setup in rodents and prevented allograft
vasculopathy in rats (Busque et al. 2009). Hence, the JAK3 inhibitor tofacitinib has shown promise
as a CNI substitute in immunosuppressive protocols. In humans, favorable nonhuman primate and
Phase I results, a Phase IIa pilot study comparing two doses of tofacitinib (15 and 30 mg b.i.d.) as
group 1 and 2 to tacrolimus as group 3 in de novo kidney allograft recipients was undertaken with
results published in 2009 (Vincenti et al. 2012). All subjects received IL-2 receptor antagonist
induction, MMF, and corticosteroids. The 6-month incidence of biopsy-proven acute rejection
(BPAR) was 5.3%, 21.1%, and 4.8% for 15 mg b.i.d. (T-15), 30 mg b.i.d. (T-30), and tacrolimus,
respectively. The cardiometabolic data for tofacitinib were mixed. The 6-month eGFR was similar
across the three groups: however, the eGFR of the 12-month extension study were 83.6, 77.6, and
73.3 mL/min for T-15, T-30, and tacrolimus, respectively. By month 12, there was no difference
between the three groups in regard to hemoglobin concentration or in the incidence of new onset
diabetes mellitus after transplant (NODAT). Compared to tacrolimus at 12 months, subjects
enrolled in the high dose tofacitinib group (T-30) experienced a significantly higher rate of BK
virus nephropathy (20%) as well as cytomegalovirus (CMV) disease (21.1%). In comparison, the 12­
month rate of BK virus nephropathy and CMV disease in tacrolimus-treated patients was 0%. The
conclusion from this study was that the preferred dose of tofacitinib was 15 mg b.i.d. as the 30 mg
b.i.d. dose regimen was associated with overimmunosuppression without improved efficacy. In the
next stage, a large Phase IIb trial was performed to evaluate the effectiveness of two different lower
dose strategies of tofacitinib compared to a CsA-based regimen. Group 1 received the active
comparator CsA. Group 2 received T-15 for months 1–6, then 10 mg b.i.d. (T-10) for months
7–12, while group 3 received T-15 for months 1–3 followed by T-10 for months 4–12. All three
groups also received MMF and steroids. The 12-month BPAR rate were 18.8%, 17.4%, and 15.4%
for groups 1, 2, and 3, respectively. Both tofacitinib-dosing strategies were non-inferior to CsA. The
12-month mean measured eGFR was significantly better for both tofacitinib groups: 53.9, 64.6, and
64.7 mL/min for groups 1, 2 and 3, respectively. The incidence of CAN in protocol biopsies was also
lower for tofacitinib-treated patients: 48.3%, 25%, and 23.9% for groups 1, 2, and 3, respectively.
The incidence of NODAT was also lower for tofacitinib-treated patients at 9.9% and 9.3% for
Renal Disorders 169

groups 2 and 3, respectively. This is in comparison to CsA with an incidence of NODAT of 20.8%.
Overall, both tofacitinib groups demonstrated noninferiority in 12-month BPAR rate and statisti­
cally higher measured eGFR compared to CsA. There was an additional benefit in tofacitinib­
treated patients of a lower rate of CAN and NODAT. It is clear that tofacitinib results in potent
immunosuppression and further tweaking of the regimen may be required to minimize the risk of
immune deficiency. Given the promising results with a large Phase IIb trial, a long-term extension
(LTE) was performed to evaluate the patients who completed 12 months of CsA or tofacitinib
treatment in the phase IIb parent study were enrolled into the LTE study. Patients were analyzed by
tofacitinib less-intensive (LI) or more-intensive (MI) regimens received in the parent study. For both
groups, tofacitinib dose was reduced from 10 to 5 mg twice daily by 6 months into the LTE.
Patients were followed up through month 72 post-transplant, with a focus on month 36 results.
Tofacitinib demonstrated similar 36-month patient and graft survival rates to CsA. Biopsy-proven
acute rejection rates at month 36 were 11.2% for CsA, versus 10.0% and 7.4% (both p > 0.05) for
tofacitinib LI and MI, respectively. Least GFR were 9 to 15 mL/min per 1.73m2 higher for
tofacitinib versus CsA at month 36. The proportions of patients with grade 2/3 interstitial fibrosis
and tubular atrophy in month 36 protocol biopsies were 20.0% for LI and 18.2% for MI (both p >
0.05) versus 33.3% for CsA. Kaplan-Meier cumulative serious infection rates at month 36 were
numerically higher for tofacitinib LI (43.9%: p = 0.45) and significantly higher for MI (55.9%: p <
0.05) versus CsA (37.1%). Long-term tofacitinib continued to be effective in preventing renal
allograft acute rejection and preserving renal function. However, long-term tofacitinib and MMF
product combination was associated with persistent serious infection risk. Further studies have been
needed to evaluate the benefit of anti-reject therapy and serious infection by the blockade of
JAK-STAT signaling pathway.

10.5 Conclusion
Current therapeutic strategy in diabetic nephropathy, lupus nephritis, and kidney transplantation
is based on the blockade of ANG II-activation by the ACE/ARB and inflammatory cytokine
cascade by the steroids and immunosuppressive agents such as cyclophosphamide, MMF, and
CNI, which confers the inefficient and the risk of the serious infection and renal toxicity. The
JAK-STAT pathway transmits signals from extracellular ligands, including many cytokines and
chemokines as well as growth factors and hormones, directly to the nucleus to induce a variety of
cellular responses. Since the JAK-STAT pathways play major activating roles in a variety of
disease processes, there has been a robust effort to develop specific inhibitors of this pathway.
JAK-STAT pathway and its components are good candidates for immunological intervention for
various disease controls.
The growing evidence for JAK-STAT activation in the pathogenesis of renal disorders estab­
lishes a new set of targets for potential intervention in this disease. Whether JAK inhibitors will
show a positive effect on the progression of renal diseases in terms of loss of kidney function and
development of ESRD and other complications remains uncertain, though the preclinical and
early clinical data are quite encouraging. However, we should consider whether the long-term
administration of JAK inhibitors in the treatment of renal diseases will be safe. These agents were
developed for the treatment of autoimmune conditions as well as myelodysplastic syndrome, and
treatment to date with any of the inhibitors has been of significantly shorter duration than the
many years of treatment that effective renal diseases therapy would be likely to entail. For
example, the long-term effect of JAK2 inhibition may be problematic if the treatment potentially
aggravates the anemia that many renal disorders patients have. Despite these concerns, there is
good reason to be optimistic about the outcomes of long-term JAK inhibition as a treatment that
could substantially slow the progression of renal disorders and make outcomes better for the
millions of individuals worldwide who suffer from this often-fatal disease.
170 JAK-STAT Signaling in Diseases

REFERENCES
Amiri, F., S. Shaw, X. Wang et al. 2002. Angiotensin II activation of the JAK/STAT pathway in
mesangial cells is altered by high glucose. Kidney Int 61(5):1605–1616.
Andrews, B. S., R. A. Eisenberg, A. N. Theofilopoulos et al. 1978. Spontaneous murine lupus-like
syndromes. Clinical and immunopathological manifestations in several strains. J Exp Med 148
(5):1198–1215.
Aringer, M., J. S. Smolen. 2004. Tumour necrosis factor and other pro-inflammatory cytokines in
systemic lupus erythematosus: a rationale for therapeutic intervention. Lupus 13(5):344–347.
Bagavant, H., S. M. Fu. 2009. Pathogenesis of kidney disease in systemic lupus erythematosus. Curr Opin
Rheumatol 21(5):489–494.
Banes, A. K., S. Shaw, J. Jenkins et al. 2004. Angiotensin II blockade prevents hyperglycemia-induced
activation of JAK and STAT proteins in diabetic rat kidney glomeruli. Am J Physiol Renal Physiol
286(4):653–659.
Baxter, E. J., L. M. Scott, P. J. Campbell et al. 2005. Acquired mutation of the tyrosine kinase JAK2 in
human myeloproliferative disorders. Lancet 365:1054–1061.
Berthier, C. C., H. Zhang, M. Schin et al. 2009. Enhanced expression of Janus kinase-signal transducer
and activator of transcription pathway members in human diabetic nephropathy. Diabetes 58
(2):469–477.
Bhat, G. J., T. J. Thekkumkara, W. G. Thomas et al. 1994. Angiotensin II stimulates sis-inducing
factor-like DNA binding activity: evidence that the AT1A receptor activates transcription
factor-Stat91 and/or a related protein. J Biol Chem 269(50):31443–31449.
Bhat, G. J., T. J. Thekkumkara, W. G. Thomas et al. 1995. Activation of the STAT pathway by
angiotensin II in T3CHO/AT1A cells: cross-talk between angiotensin II and interleukin-6 nuclear
signaling. J Biol Chem 270(32):19059–19065.
Blanco, P., V. Pitard, J. F. Viallard et al. 2005. Increase in activated CD8+ T lymphocytes expressing
perforin and granzyme B correlates with disease activity in patients with systemic lupus
erythematosus. Arthritis Rheum 52(1):201–211.
Boers, M., M. T. Nurmohamed, C. J. Doelman et al. 2003. Influence of glucocorticoids and disease
activity on total and high density lipoprotein cholesterol in patients with rheumatoid arthritis. Ann
Rheum Dis 62(9):842–845.
Boggio, E., N. Clemente, A. Mondino et al. 2004. IL-17 protects T cells from apoptosis and contributes to
development of ALPS-like phenotypes. Blood 123(8):1178–1186.
Borie, D. C., P. S. Changelian, M. J. Larson et al. 2005. Immunosuppression by the JAK3 inhibitor
CP-690,550 delays rejection and significantly prolongs kidney allograft survival in nonhuman
primates. Transplantation 79(7):791–801.
Borie, D. C., M. S. Si, R. E. Morris et al. 2003. JAK3 inhibition as a new concept for immune suppression.
Curr Opin Investig Drugs 4(11):1297–1303.
Boyle, D. L., K. Soma, J. Hodge et al. 2015. The JAK inhibitor tofacitinib suppresses synovial
JAK1-STAT signalling in rheumatoid arthritis. Ann Rheum Dis 74(6):1311–1316.
Brosius, F. C., 3rd, J. C. He. 2015. JAK inhibition and progressive kidney disease. Curr Opin Nephrol
Hypertens 24(1):88–95.
Brosius, F. C., K. R. Tuttle, M. Kretzler. 2016. JAK inhibition in the treatment of diabetic kidney disease.
Diabetologia 59(8):1624–1627.
Buckley, R. H. 2004. The multiple causes of human SCID. J Clin Invest 114(10):1409–1411.
Burmester, G. R., J. M. Kremer, F. Van Den Bosch et al. 2018. Safety and efficacy of upadacitinib in
patients with rheumatoid arthritis and inadequate response to conventional synthetic
disease-modifying anti-rheumatic drugs (SELECT-NEXT): a randomised, double-blind,
placebo-controlled phase 3 trial. Lancet 391(10139):2503–2512.
Busque, S., J. Leventhal, D. C. Brennan et al. 2009. Calcineurin-inhibitor-free immunosuppression based
on the JAK inhibitor CP-690,550: a pilot study in de novo kidney allograft recipients. Am
J Transplant 9(8):1936–1945.
Cash, H., M. Relle, J. Menke et al. 2010. Interleukin 6 (IL-6) deficiency delays lupus nephritis in
MRL-Faslpr mice: the IL-6 pathway as a new therapeutic target in treatment of autoimmune
kidney disease in systemic lupus erythematosus. J Rheumatol 37(1):60–70.
Renal Disorders 171

Changelian, P. S., M. E. Flanagan, D. J. Ball et al. 2003. Prevention of organ allograft rejection by
a specific Janus kinase 3 inhibitor. Science 302(5646):875–878.
Chappey, O., C. Dosquet, M. P. Wautier et al. 1997. Advanced glycation end products, oxidant stress and
vascular lesions. Eur J Clin Invest 27(2):97–108.
Charles-Schoeman, C., R. Fleischmann, J. Davignon et al. 2015. Potential mechanisms leading to the
abnormal lipid profile in patients with rheumatoid arthritis versus healthy volunteers and reversal
by tofacitinib. Arthritis Rheumatol 67(3):616–625.
Chen, J., J. K. Chen, R. C. Harris. 2012. Angiotensin II induces epithelial-to-mesenchymal transition in
renal epithelial cells through reactive oxygen species/Src/caveolin-mediated activation of an epider­
mal growth factor receptor-extracellular signal-regulated kinase signaling pathway. Mol Cell Biol 32
(5):981–991.
Choudhury, G. G., N. Ghosh-Choudhury, H. E. Abboud. 1998. Association and direct activation of
signal transducer and activator of transcription1alpha by platelet-derived growth factor receptor.
J Clin Invest 101(12):2751–2760.
Chun, H. Y., J. W. Chung, H. A. Kim et al. 2007. Cytokine IL-6 and IL-10 as biomarkers in systemic
lupus erythematosus. J Clin Immunol 27(5):461–466.
Clark, J. D., M. E. Flanagan, J. B. Telliez. 2014. Discovery and development of Janus kinase (JAK)
inhibitors for inflammatory diseases. J Med Chem 57(12):5023–5038.
Cohen, P. L., R. A. Eisenberg. 1991. Lpr and gld: single gene models of systemic autoimmunity and
lymphoproliferative disease. Annu Rev Immunol 9:243–269.
Conklyn, M., C. Andresen, P. Changelian et al. 2004. The JAK3 inhibitor CP-690550 selectively reduces
NK and CD8+ cell numbers in cynomolgus monkey blood following chronic oral dosing. J Leukoc
Biol 76(6):1248–1255.
Craiglow, B. G., B. A. King. 2015. Tofacitinib citrate for the treatment of vitiligo: a pathogenesis-directed
therapy. JAMA Dermatol 151(10):1110–1112.
Crispin, J. C., M. Oukka, G. Bayliss et al. 2008. Expanded double negative T cells in patients with systemic
lupus erythematosus produce IL-17 and infiltrate the kidneys. J Immunol 181(12):8761–8766.
Croker, B. A., H. Kiu, S. E. Nicholson. 2008. SOCS regulation of the JAK/STAT signaling pathway.
Semin Cell Dev Biol 19(4):414–422.
Denny, M. F., S. Yalavarthi, W. Zhao et al. 2010. A distinct subset of pro-inflammatory neutrophils
isolated from patients with systemic lupus erythematosus induces vascular damage and synthesizes
type I IFNs. J Immunol 184(6):3284–3297.
Ducloux, D., G. Motte, M. Kribs et al. 2002. Hypertension in renal transplantation: donor and recipient
risk factors. Clin Nephrol 57(6):409–413.
Edgerton, C., J. C. Crispin, C. M. Moratz et al. 2009. IL-17 producing CD4+ T cells mediate
accelerated ischemia/reperfusion-induced injury in autoimmunity-prone mice. Clin Immunol
130(3):313–321.
Febbraio, M. A., S. Rose-John, B. K. Pedersen. 2010. Is interleukin-6 receptor blockade the Holy Grail
for inflammatory diseases? Clin Pharmacol Ther 87(4):396–398.
Fridman, J. S., P. A. Scherle, R. Collins et al. 2010. Selective inhibition of JAK1 and JAK2 is efficacious in
rodent models of arthritis: preclinical characterization of INCB028050. J Immunol 184(9):5298–5307.
Fujimoto, M., H. Tsutsui, O. Xinshou et al. 2004. Inadequate induction of suppressor of cytokine
signaling-1 causes systemic autoimmune diseases. Int Immunol 16(2):303–314.
Furumoto, Y., M. Gadina. 2013. The arrival of JAK inhibitors: advancing the treatment of immune and
hematologic disorders. Bio Drugs 27(5):431–438.
Furumoto, Y., C. K. Smith, L. Blanco et al. 2017. Tofacitinib ameliorates murine lupus and its associated
vascular dysfunction. Arthritis Rheumatol 69(1):148–160.
Gaston, R. S., J. M. Cecka, B. L. Kasiske et al. 2010. Evidence for antibody-mediated injury as a major
determinant of late kidney allograft failure. Transplantation 90(1):68–74.
Genovese, M. C., M. Greenwald, C. Codding et al. 2017. Peficitinib, a JAK inhibitor, in combination with
limited conventional synthetic disease-modifying antirheumatic drugs in the treatment of
moderate-to-severe rheumatoid arthritis. Arthritis Rheumatol 69(5):932–942.
Genovese, M. C., F. Yang, M. Østergaard et al. 2016. Efficacy of VX-509 (decernotinib) in combination
with a disease-modifying antirheumatic drug in patients with rheumatoid arthritis: clinical and
MRI findings. Ann Rheum Dis 75(11):1979–1983.
172 JAK-STAT Signaling in Diseases

Ghoreschi, K., M. I. Jesson, X. Li et al. 2011. Modulation of innate and adaptive immune responses by
tofacitinib (CP-690,550). J Immunol 186(7):4234–4243.
Ghoreschi, K., A. Laurence, J. J. O’Shea. 2009. Janus kinases in immune cell signaling. Immunol Rev 228
(1):273–287.
Gohda, T., M. A. Niewczas, L. H. Ficociello et al. 2012. Circulating TNF receptors 1 and 2 predict stage 3
CKD in type 1 diabetes. J Am Soc Nephrol 23(3):516–524.
Grisouard, J., H. Hao-Shen, S. Dirnhofer et al. 2014. Selective deletion of Jak2 in adult mouse
hematopoietic cells leads to lethal anemia and thrombocytopenia. Haematologica 99(4):52–54.
Ha, H., H. B. Lee. 2000. Reactive oxygen species as glucose signaling molecules in mesangial cells
cultured under high glucose. Kidney Int Supp l77:S19–S25.
Harrison, C., J. J. Kiladjian, H. K. Al-Ali et al. 2012. JAK inhibition with ruxolitinib versus best available
therapy for myelofibrosis. N Engl J Med 366(9):787–798.
Holland, S. M., F. R. DeLeo, H. Z. Elloumi. 2007. STAT3 mutations in the hyper-IgE syndrome. N Engl
J Med 357(16):1608–1619.
Illei, G. G., Y. Shirota, C. H. Yarboro et al. 2010. Tocilizumab in systemic lupus erythematosus: data on
safety. preliminary efficacy, and impact on circulating plasma cells from an open-label phase I
dosage-escalation study. Arthritis Rheum 62(2):542–552.
James, C., V. Ugo, J. P. Le Couédic et al. 2005. A unique clonal JAK2 mutation leading to constitutive
signalling causes polycythaemia vera. Nature 434(7037):1144–1148.
Jedidi, A., C. Marty, C. Oligo et al. 2009. Selective reduction of JAK2V617F-dependent cell growth by
siRNA/shRNA and its reversal by cytokines. Blood 114(9):1842–1851.
Jeon, J. Y., K. Y. Kim, H. A. Kim et al. 2010. Interleukin 6 gene polymorphisms are associated with
systemic lupus erythematosus in Koreans. J Rheumatol 37(11):2251–2258.
Jevnikar, A. M., R. B. Mannon. 2008. Late kidney allograft loss: what we know about it and what we can
do about it. Clin J Am Soc Nephrol 3(Suppl 2):S56–567.
Johnson, R. J., J. Floege, W. G. Couser et al. 1993. Role of platelet-derived growth factor in glomerular
disease. J Am Soc Nephrol 4(2):119–128.
Kaplan, M. J. 2011. Neutrophils in the pathogenesis and manifestations of SLE. Nat Rev Rheumatol 7
(12):691–699.
Karaman, M. W., S. Herrgard, D. K. Treiber et al. 2008. A quantitative analysis of kinase inhibitor
selectivity. Nat Biotechnol 26(1):127–132.
Kavanaugh, A., J. Kremer, L. Ponce et al. 2017. Filgotinib (GLPG0634/GS-6034), an oral selective JAK1
inhibitor, is effective as monotherapy in patients with active rheumatoid arthritis: results from
a randomised, dose-finding study (DARWIN 2). Ann Rheum Dis 76(6):1009–1019.
Kirou, K. A., E. Gkrouzman. 2013. Anti-interferon alpha treatment in SLE. Clin Immunol 148(3):303–312.
Knight, J. S., W. Zhao, W. Luo et al. 2013. Peptidylarginine deiminase inhibition is immunomodulatory
and vasculoprotective in murine lupus. J Clin Invest 123(7):2981–2993.
Kralovics, R., F. Passamonti, A. S. Buser et al. 2005. A gain-of-function mutation of JAK2 in
myeloproliferative disorders. N Engl J Med 352(17):1779–1790.
Krempler, A., Y. Qi, A. A. Triplett et al. 2004. Generation of a conditional knockout allele for the Janus
kinase 2 (Jak2) gene in mice. Genesis 40(1):52–57.
Kudlacz, E., B. Perry, P. Sawyer et al. 2004. The novel JAK–3 inhibitor CP-690550 is a potent
immunosuppressive agent in various murine models. Am J Transplant 4(1):51–57.
Kuhrt, D., D. M. Wojchowski. 2015. Emerging EPO and EPO receptor regulators and signal transducers.
Blood 125(23):3536–3541.
Lan, H. Y., A. C. Chung. 2012. TGF-β/Smad signaling in kidney disease. Semin Nephrol 32(3):236–243.
Lannutti, B. J., A. Epp, J. Roy et al. 2009. Incomplete restoration of Mpl expression in the mpl-/- mouse
produces partial correction of the stem cell-repopulating defect and paradoxical thrombocytosis.
Blood 113(8):1778–1785.
Levine, R. L., M. Wadleigh, J. Cools et al. 2005. Activating mutation in the tyrosine kinase JAK2 in
polycythemia vera, essential thrombocythemia, and myeloid metaplasia with myelofibrosis. Cancer
Cell 7(4):387–397.
Levy, L. L., J. Urban, B. A. King. 2015. Treatment of recalcitrant atopic dermatitis with the oral Janus
kinase inhibitor tofacitinib citrate. J Am Acad Dermatol 73(3):395–399.
Renal Disorders 173

Li, J., G. Gobe. 2006. Protein kinase C activation and its role in kidney disease. Nephrology (Carlton) 11
(5):428–434.
Li, X., Y. Wang, K. Wang, Y. Wu et al. 2018. Renal protective effect of Paeoniflorin by inhibition of
JAK2/STAT3 signaling pathway in diabeticmice. Biosci Trends 12(2):168–176.
Ling, B. N., E. E. Seal, D. C. Eaton. 1993. Regulation of mesangial cell ion channels by insulin and
angiotensin II. Possible role in diabetic glomerular hyperfiltration. J Clin Invest 92(5):2141–2151.
Liu, R., Y. Zhong, X. Li et al. 2014. Role of transcription factor acetylation in diabetic kidney disease.
Diabetes 63(7):2440–2453.
Lodhi, S. A., K. E. Lamb, H. U. Meier-Kriesche. 2011. Solid organ allograft survival improvement in the
United States: the long-term does not mirror the dramatic short-term success. Am J Transplant 11
(6):1226–1235.
López-Pedrera, C., M. A. Aguirre, N. Barbarroja et al. 2010. Accelerated atherosclerosis in systemic lupus
erythematosus: role of pro-inflammatory cytokines and therapeutic approaches. J Biomed Biotech­
nol 2010:607084.
Lu, T. C., Z. H. Wang, X. Feng et al. 2009. Knockdown of Stat3 activity in vivo prevents diabetic
glomerulopathy. Kidney Int 76(1):63–71.
Mahajan, S., J. K. Hogan, D. Shlyakhter et al. 2015. VX-509 (decernotinib) is a potent and selective janus
kinase 3 inhibitor that attenuates inflammation in animal models of autoimmune disease.
J Pharmacol Exp Ther 353(2):405–414.
Manshouri, T., A. Quintás-Cardama, R. H. Nussenzveig et al. 2008. The JAK kinase inhibitor
CP-690,550 suppresses the growth of human polycythemia vera cells carrying the JAK2V617F
mutation. Cancer Science 99(6):1265–1273.
Markham, A. 2017. Baricitinib: first global approval. Drugs 77(6):697–704.
Marrero, M. B., A. K. Banes-Berceli, D. M. Stern et al. 2006. Role of the JAK/STAT signaling pathway in
diabetic nephropathy. Am J Physiol Renal Physiol 290(4):F762–7768.
Mathis, A. S., N. Dave, G. T. Knipp et al. 2004. Drug-related dyslipidemia after renal transplantation. Am
J Health Syst Pharm 61(6):565–585.
Mazière, C., M. A. Conte, J. C. Mazière. 2001. Activation of JAK2 by the oxidative stress generated with
oxidized low-density lipoprotein. Free Radic Biol Med 31(11):1334–1340.
McInnes, I. B., R. Higgs, J. Lee. 2017. Ex vivo comparison of baricitinib, upadacitinib, filgotinib, and
tofacitinib for cytokine signaling in human leukocyte subpopulations. Arthritis Rheumatol 69(Suppl
10):Abstract 2870.
Meier-Kriesche, H. U., J. D. Schold, T. R. Srinivas et al. 2004. Lack of improvement in renal allograft
survival despite a marked decrease in acute rejection rates over the most recent era. Am J Transplant
4(3):378–383.
Mesa, R. A., J. Cortes. 2013. Optimizing management of ruxolitinib in patients with myelofibrosis: the
need for individualized dosing. J Hematol Oncol 6:79.
Migita, K., Y. Izumi, T. Torigoshi et al. 2013. Inhibition of Janus kinase/signal transducer and activator of
transcription (JAK/STAT) signaling pathway in rheumatoid synovial fibroblasts using small
molecule compounds. Clin Exp Immunol 174(3):356–363.
Mohamed, M. F., H. S. Camp, P. Jiang et al. 2016. Pharmacokinetics, safety and tolerability of ABT-494,
a novel selective JAK 1 inhibitor, in healthy volunteers and subjects with rheumatoid arthritis. Clin
Pharmacokinet 55(12):1547–1558.
Navarro-González, J. F., C. Mora-Fernández, M. Muros de Fuentes et al. 2011. Inflammatory molecules
and pathways in the pathogenesis of diabetic nephropathy. Nat Rev Nephrol 7(6):327–340.
Neubauer, H., A. Cumano, M. Müller et al. 1998. Jak2 deficiency defines an essential developmental
checkpoint in definitive hematopoiesis. Cell 93(3):397–409.
Ng, A. P., M. Kauppi, D. Metcalf et al. 2014. Mpl expression on megakaryocytes and platelets is
dispensable for thrombopoiesis but essential to prevent myeloproliferation. Proc Natl Acad Sci
U S A 111(16):5884–5889.
Niewczas, M. A., T. Gohda, J. Skupien et al. 2012. Circulating TNF receptors 1 and 2 predict ESRD in
type 2 diabetes. J Am Soc Nephrol 23(3):507–515.
Niewold, T. B., J. Hua, T. J. Lehman et al. 2007. High serum IFN-alpha activity is a heritable risk factor
for systemic lupus erythematosus. Genes Immun 8(6):492–502.
174 JAK-STAT Signaling in Diseases

O’Shea, J. J., R. Plenge. 2012. JAK and STAT signaling molecules in immunoregulation and
immune-mediated disease. Immunity 36(4):542–550.
O’Shea, J. J., D. M. Schwartz, A. V. Villarino et al. 2015. The JAK–STAT pathway: impact on human
disease and therapeutic intervention. Annu Rev Med 66:311–328.
Okiyama, N., Y. Furumoto, V. A. Villarroel et al. 2014. Reversal of CD8 T-cell-mediated mucocutaneous
graft-versus-host-like disease by the JAK inhibitor tofacitinib. J Invest Dermatol 134(4):992–1000.
Olivera, P., S. Danese, L. Peyrin-Biroulet. 2017. JAK inhibition in inflammatory bowel disease. Expert
Rev Clin Immunol 13(7):693–703.
Õrtiz-Munoz, G., V. Lopez-Parra, O. Lopez-Franco et al. 2010. Suppressors of cytokine signaling
abrogate diabetic nephropathy. J Am Soc Nephrol 21(5):763–772.
Paniagua, R., M. S. Si, M. G. Flores et al. 2005. Effects of JAK3 inhibition with CP-690,550 on immune
cell populations and their functions in nonhuman primate recipients of kidney allografts. Trans­
plantation 80(9):1283–1292.
Parganas, E., D. Wang, D. Stravopodis et al. 1998. Jak2 is essential for signaling through a variety of
cytokine receptors. Cell 93(3):385–395.
Park, S. O., H. L. Wamsley, K. Bae et al. 2013. Conditional deletion of Jak2 reveals an essential role in
hematopoiesis throughout mouse ontogeny: implications for Jak2 inhibition in humans. PLoS One
8(3):e59675.
Perry, D., A. Sang, Y. Yin et al. 2011. Murine models of systemic lupus erythematosus. J Biomed
Biotechnol 2011:271694.
Pikman, Y., B. H. Lee, T. Mercher et al. 2006. MPLW515L is a novel somatic activating mutation in
myelofibrosis with myeloid metaplasia. PLoS Med 3(7):270.
Podder, H., B. D. Kahan. 2004. Janus kinase 3: a novel target for selective transplant immunosupression.
Expert Opin Ther Targets 8(6):613–629.
Quintás-Cardama, A., K. Vaddi, P. Liu, T. Manshouri et al. 2010. Preclinical characterization of the
selective JAK1/2 inhibitor INCB018424: therapeutic implications for the treatment of myeloproli­
ferative neoplasms. Blood 115(15):3109–3117.
Richez, C., M. E. Truchetet, M. Kostine et al. 2017. Efficacy of baricitinib in the treatment of rheumatoid
arthritis. Expert Opin Pharmacother 18(13):1399–1407.
Ripley, B. J., B. Goncalves, D. A. Isenberg et al. 2005. Raised levels of interleukin 6 in systemic lupus
erythematosus correlate with anaemia. Ann Rheum Dis 64(6):849–853.
Rousvoal, G., M. S. Si, M. Lau et al. 2006. Janus kinase 3 inhibition with CP-690,550 prevents allograft
vasculopathy. Transpl Int 19(12):1014–1021.
Saran, R., Y. Li, B. Robinson, J. Ayanian et al. 2015. US Renal Data System 2014 annual data report:
epidemiology of kidney disease in the United States. Am J Kidney Dis 66(1 Suppl):1–305.
Schwartz, D. M., M. Bonelli, M. Gadina. 2016. Type I/II cytokines. JAKs, and new strategies for treating
autoimmune diseases. Nat Rev Rheumatol 12(1):25–36.
Scott, L. M., W. Tong, R. L. Levine et al. 2007. JAK2 exon 12 mutations in polycythemia vera and
idiopathic erythrocytosis. N Engl J Med 356(5):459–468.
Shaw, S., X. D. Wang, H. Redd et al. 2003. High glucose augments the angiotensin II-induced
activation of JAK2 in vascular smooth muscle cells via the polyol pathway. J Biol Chem 278
(33):30634–30641.
Shi, J. G., X. Chen, F. Lee et al. 2014. The pharmacokinetics, pharmacodynamics, and safety of
baricitinib, an oral JAK 1/2 inhibitor, in healthy volunteers. J Clin Pharmacol 54(12):1354–1361.
Silver, J. S., C. A. Hunter. 2010. gp130 at the nexus of inflammation, autoimmunity, and cancer. J Leukoc
Biol 88(6):1145–1156.
Simon, A. R., U. Rai, B. L. Fanburg et al. 1998. Activation of the JAK–STAT pathway by reactive oxygen
species. Am J Physiol Cell Physiol 275(6):1640–1652.
Siragy, H. M., R. M. Carey. 2010. Role of the intrarenal renin-angiotensin-aldosterone system in chronic
kidney disease. Am J Nephrol 31(6):541–550.
Tam, C. S., S. Verstovsek. 2013. Investigational Janus kinase inhibitors. Expert Opin Investig Drugs 22
(6):687–699.
Thacker, S. G., C. C. Berthier, D. Mattinzoli et al. 2010. The detrimental effects of IFN-alpha on
vasculogenesis in lupus are mediated by repression of IL-1 pathways: potential role in atherogenesis
and renal vascular rarefaction. J Immunol 185(7):4457–4469.
Renal Disorders 175

Thacker, S. G., X. Rousset, S. Esmail et al. 2015. Increased plasma cholesterol esterification by LCAT reduces
diet-induced atherosclerosis in SR-BI knockout mice. Journal of Lipid Research 56(7):1282–1295.
Thacker, S. G., W. Zhao, C. K. Smith et al. 2012. Type I interferons modulate vascular function, repair,
thrombosis, and plaque progression in murine models of lupus and atherosclerosis. Arthritis Rheum
64(9):2975–2985.
Tiedt, R., J. Coers, S. Ziegler et al. 2009. Pronounced thrombocytosis in transgenic mice expressing reduced
levels of Mpl in platelets and terminally differentiated megakaryocytes. Blood 113(8):1768–1777.
Townsend, M. J., J. G. Monroe, A. C. Chan. 2010. B-cell targeted therapies in human autoimmune
diseases: an updated perspective. Immunol Rev 237(1):264–283.
Tsokos, G. C. 2011. Systemic lupus erythematosus. N Engl J Med 365(22):2110–2121.
Tucci, M., L. Lombardi, H. B. Richards et al. 2008. Overexpression of interleukin-12 and T helper 1
predominance in lupus nephritis. Clin Exp Immunol 154(2):247–254.
Tuttle, K. R., F. C. Brosius, 3rd, S. G. Adler et al. 2018. JAK1/JAK2 inhibition by baricitinib in diabetic
kidney disease: results from a Phase 2 randomized controlled clinical trial. Nephrol Dial Transplant
33(11):1950–1959.
Ushio-Fukai, M., R. W. Alexander, M. Akers et al. 1999. Reactive oxygen species mediate the activation
of Akt/protein kinase B by angiotensin II in vascular smooth muscle cells. J Biol Chem 274
(32):22699–22704.
Usui, F., H. Kimura, T. Ohshiro et al. 2012. Interleukin-17 deficiency reduced vascular inflammation and
development of atherosclerosis in Western diet-induced apoE-deficient mice. Biochemical Biophys
Res Commun 420(1):72–77.
Van, R. L., R. Galien, E. M. van der Aar et al. 2013. Preclinical characterization of GLPG0634, a selective
inhibitor of JAK1, for the treatment of inflammatory diseases. J Immunol 191(7):3568–3577.
Verstovsek, S., H. Kantarjian, R. A. Mesa et al. 2010. Safety and efficacy of INCB018424, a JAK1 and
JAK2 inhibitor, in myelofibrosis. N Engl J Med 363(12):1117–1127.
Verstovsek, S., R. A. Mesa, J. Gotlib et al. 2012. A double-blind, placebo-controlled trial of ruxolitinib
for myelofibrosis. N Engl J Med 366(9):799–807.
Vincenti, F., S. Friman, E. Scheuermann et al. 2007. Results of an international, randomized trial
comparing glucose metabolism disorders and outcome with cyclosporine versus tacrolimus. Am
J Transplant 7(6):1506–1514.
Vincenti, F., A. D. Kirk. 2008. What’s next in the pipeline. Am J Transplant 8(10):1972–1981.
Vincenti, F., S. H. Tedesco, S. Busque et al. 2012. Randomized phase 2b trial of tofacitinib (CP-690,550)
in de novo kidney transplant patients: efficacy, renal function and safety at 1 year. Am J Transplant
12(9):2446–2456.
Wada, J., H. Makino. 2013. Inflammation and the pathogenesis of diabetic nephropathy. Clin Sci (Lond)
124(3):139–152.
Wang, S., N. Yang, L. Zhang et al. 2010. Jak/STAT signaling is involved in the inflammatory infiltration
of the kidneys in MRL/lpr mice. Lupus 19(10):1171–1180.
Wang, X., S. Shaw, F. Amiri et al. 2002. Inhibition of the Jak/STAT signaling pathway prevents the high
glucose-induced increase in TGF-beta and fibronectin synthesis in mesangial cells. Diabetes 51
(12):3505–3509.
Wenzel, U. O., B. Fouqueray, P. Biswas et al. 1995. Activation of mesangial cells by the phosphatase
inhibitor vanadate. Potential implications for diabetic nephropathy. J Clin Invest 95(3):1244–1252.
Westhovens, R., P. C. Taylor, R. Alten et al. 2017. Filgotinib (GLPG0634/GS-6034), an oral JAK1
selective inhibitor, is effective in combination with methotrexate (MTX) in patients with active
rheumatoid arthritis and insufficient response to MTX: results from a randomised, dose-finding
study (DARWIN 1). Ann Rheum Dis 76(6):998–1008.
Wiik, A. C., C. Wade, T. Biagi et al. 1994. A deletion in nephronophthisis 4 (NPHP4) is associated with
recessive cone-rod dystrophy in standard wire-haired dachshund. Genome Res 18(9):1415–1421.
Winter, O., K. Moser, E. Mohr et al. 2010. Megakaryocytes constitute a functional component of
a plasma cell niche in the bone marrow. Blood 116(11):1867–1875.
Witthuhn, B. A., F. W. Quelle, O. Silvennoinen et al. 1993. JAK2 associates with the erythropoietin
receptor and is tyrosine phosphorylated and activated following stimulation with erythropoietin.
Cell 74(2):227–236.
176 JAK-STAT Signaling in Diseases

Witthuhn, B. A., O. Silvennoinen, O. Miura et al. 1994. Involvement of the JAK–3 Janus kinase in
signalling by interleukins 2 and 4 in lymphoid and myeloid cells. Nature 370(6485):153–157.
Wolfe, R. A., V. B. Ashby, E. L. Milford et al. 1999. Comparison of mortality in all patients on dialysis,
patients on dialysis awaiting transplantation, and recipients of a first cadaveric transplant. N Engl
J Med 341(23):1725–1730.
Wong, C. K., L. C. Lit, L. S. Tam et al. 2008. Hyperproduction of IL-23 and IL-17 in patients with
systemic lupus erythematosus: implications for Th17-mediated inflammation in auto-immunity.
Clin Immunol 127(3):385–393.
Woroniecka, K. I., A. S. Park, D. Mohtat et al. 2011. Transcriptome analysis of human diabetic kidney
disease. Diabetes 60(9):2354–2369.
Xing, L., Z. Dai, A. Jabbari et al. 2014. Alopecia areata is driven by cytotoxic T lymphocytes and is
reversed by JAK inhibition. Nature Med 20(9):1043–1049.
Yu, C. C., A. A. Mamchak, A. L. DeFranco. 2003. Signaling mutations and autoimmunity. Curr Dir
Autoimmun 6:61–88.
Zhang, H., V. Nair, J. Saha, K. B. Atkins et al. 2017. Podocyte-specific JAK2 overexpression worsens
diabetic kidney disease in mice. Kidney Int 92(4):909–921.
Zhuang, S. 2013. Regulation of STAT signaling by acetylation. Cell Signal 25(9):1924–1931.
11
JAK-STAT Signaling in Hematologic Malignancies

Thomas Pincez
Division of Pediatric Hematology-Oncology
Charles-Bruneau Cancer Center, Ste-Justine Hospital
Quebec, Canada

Thai Hoa Tran


Division of Pediatric Hematology-Oncology
Charles-Bruneau Cancer Center, Ste-Justine Hospital
Quebec, Canada
Department of Pediatrics
University of Montreal
Quebec, Canada

11.1 Introduction
Hematopoiesis leads to the production of different types of mature blood cells derived from
hematopoietic stem cells (HSC). This process is highly regulated by cytokines, which govern cell fate
choices by inducing intracellular signals. Throughout hematopoiesis, cytokine signaling is under tight
control to regulate cellular proliferation and differentiation. As most cytokine receptors lack intrinsic
kinase activity, they are coupled to other signaling intermediates to mediate signal transduction. JAK­
STAT proteins are a major system of downstream signaling in hematopoietic ontogeny. In physiology,
its critical role has been demonstrated in mice models and the loss-of-function of JAK1, JAK3,
TYK2, STAT1, STAT2, STAT4, or STAT6 impair lymphopoiesis (Khwaja 2006). Moreover, JAK2-,
STAT3-, and STAT5-deficient mice are not viable (Khwaja 2006). In humans, several primary
immunodeficiencies resulting from JAK-STAT deficiency have been described (Bousfiha et al. 2018).
Hematological malignancies encompass neoplasms arising from blood cells at any differen­
tiation stage (Figure 11.1). The genomic aberrations underlying the malignant transformation
vary upon different entities. Some diseases share disrupted signaling pathways; but the
molecular mechanisms, functional consequences, and therapeutic relevance may differ. There­
fore, the JAK-STAT pathway alterations within each hematological malignancy will be dis­
cussed separately in this chapter. Three major categories of hematological malignancies harbor
recurrent alterations of this pathway: acute leukemias, myeloproliferative neoplasms (MPN),
and lymphomas. The 2016 revised classification of the World Health Organization will be used
in this chapter (Arber et al. 2016; Swerdlow et al. 2016).
Aside from acute lymphoblastic leukemia (ALL), the remaining diseases are more frequently
encountered in adults than children. Consequently, numerous studies are performed on adult samples
and the knowledge has to be extrapolated from these data. However, the clinical presentation,
evolution, and the underpinning biology of each malignancy may differ and data interpretation
needs to be nuanced. The relevant studies from the adult population will be discussed here and
adjusted to pediatric patients.

177
178 JAK-STAT Signaling in Diseases

FIGURE 11.1 Schematic representation of classification of hematological malignancies.


Classification according to the World Health Organization 2016 revision. In blue, the malignancies described in this
chapter with their subtypes. In grey, the other hematological malignancies.

JAK-STAT pathway involvement in several diseases led to the development of several pharmaco­
logic inhibitors, which can be divided in two categories (Vainchenker et al. 2018). Type I inhibitors
target the ATP-binding site of the JAK proteins under the active conformation of the kinase domain.
Type II inhibitors bind to the ATP-binding pocket of kinase domains in inactive conformation and
provide not only stronger inhibition, but also more side effects than type I drugs. The preclinical and
clinical data of both will be discussed for each malignancy.

11.2 Acute Leukemias


Acute leukemia is a malignant hematopoietic proliferation of immature blood cells from the bone
marrow. The involved lineage can be lymphoid or myeloid, defining ALL or acute myeloid
leukemia (AML).
ALL is markedly more frequent in the pediatric than adult population. In the Unites States of
America (USA), 35.9 cases per million of person younger than 20 years old are diagnosed
each year with ALL (Noone et al. 2018). The peak incidence lies from 3–5 years of age with
a slight male predominance (55%) (Noone et al. 2018). Over the past decades, long-term survival
rates have reached nearly 90% with modern chemotherapy regimens (Hunger and Mullighan
2015a). ALL may arise from B- or T-cell progenitors; hence called B-ALL and T-ALL,
respectively. Advances in genomic profiling and next-generation sequencing (NGS) technologies
have provided new insights into biology and defined the genomic landscape of these different
leukemia subtypes. Mature B-ALL (Burkitt leukemia) is a specific entity with distinct treatment
and prognosis. The JAK-STAT pathway has not been shown to be disrupted and this leukemia
will not be discussed here. B-ALL can develop from different stages of maturation of B-cell
JAK-STAT Signaling in Hematologic Malignancies 179

precursors and harbors recurrent genomic alterations associated with prognosis. High hyperdi­
ploidy (>50 chromosomes) and the t(12;21) translocation encoding ETV6-RUNX1, representing
each ~25% of B-ALL, are associated with favorable outcome (Hunger and Mullighan 2015a).
KMT2A rearrangements (occurring mainly in children <1 year) and hypodiploidy (<44 chromo­
somes) are both poor prognostic factors. The t(9;22), also named Philadelphia chromosome,
encodes the constitutively activated tyrosine kinase BCR-ABL1 and is present in 3–5% of
children with ALL (the incidence increases with age). Philadelphia chromosome-positive ALL
(Ph+ ALL) was historically associated with poor prognosis but targeted therapy in combination
with intensive chemotherapy backbone drastically improved survival rates (Aricò et al. 2010;
Biondi et al. 2012; Schultz et al. 2014; Slayton et al. 2018). T-ALL used to be an adverse
prognostic factor, especially in the case of early precursor T-ALL. However, outcome differences
between B-ALL and T-ALL become increasingly smaller in the era of risk-stratified intensive
chemotherapy regimens (Winter et al. 2018). In addition, recurrent cytogenetic features do not
confer prognosis and are not incorporated in the risk stratification of T-ALL treatment. In this
chapter, B- and T-ALL will be discussed separately. The relative high frequency of ALL in
children allows an in-depth investigation from exclusively pediatric samples.
The annual incidence of AML is much lower in pediatric (<20 years) than in the adult
population, 8.7 cases compared to 42.6 cases per million in the USA, respectively (Noone
et al. 2018). The highest incidence in childhood AML is found before the first year of life (18.8
cases per million), while the lowest occurs among children between 5 and 9 years old. AML
remains a clinical challenge with lower cure rates than ALL and an overall survival around 65%
despite modern chemotherapy regimens (Cooper et al. 2012; Gamis et al. 2013, 2014). If AML
classification is heavily based on morphology, its prognosis is, however, greatly influenced by the
underlying biology (Bolouri et al. 2018). A great diversity of cytogenetic and/or genomic
alterations is encountered in pediatric AML and varies according to age. Moreover, in some
cases, such as KMT2A rearrangements, the prognostic significance of a given gene rearrange­
ment depends on the involved gene partner (Creutzig et al. 2012). To allow for a clinically
relevant stratification in this heterogeneous group, subtypes of AML are defined by distinct
genomic alterations, which confer prognosis. The presence of t(8;21), inv(16), or t(16;16) are
classically favorable prognostic factors. Conversely, monosomy 7, monosomy 5/5q deletion, and
high allelic ratio FLT3 internal tandem duplication are associated with poor outcome (Rooij
et al. 2015). The genomic landscape of AML differs greatly between adults and children;
therefore, the clinical relevance of the JAK-STAT pathway within the pediatric AML popula­
tion will be highlighted in this chapter.

11.2.1 JAK-STAT Pathway Activation Is a Recurrent Feature of Ph-Like and


Down-Syndrome Associated B-ALL
11.2.1.1 Ph-like ALL
In 2009, genome-wide profiling studies identified a new subtype of B-ALL that exhibits a gene
expression signature similar to those found in Ph+ ALL but lacking the canonical BCR-ABL1 gene
fusion (Den Boer et al. 2009; Mullighan et al. 2009). Thus, this new subtype is known as BCR-
ABL1-like or Ph-like ALL, and represents approximately 20% of B-ALL. In contrast to Ph+ ALL,
in which the prevalence augments with increasing age, Ph-like ALL occurs in 14% among children
with National Cancer Institute (NCI) standard-risk B-ALL between 1 and 9 years of age, 20% in
NCI high-risk B-ALL between 10 and 21 years old, peaks at 28% among young adults between 21
and 39 years old, and decreases to 22% in older adults from 40 to 86 years old (Roberts et al. 2014,
2016, 2018). The Ph-like signature is associated with high frequency of IKZF1 alterations and
confers poor outcomes despite intensive chemotherapy regimens (Jain et al. 2017; Roberts et al.
2014; Tran et al. 2018). Ph-like ALL is particularly prevalent in patients of Hispanic ethnicity
(Harvey et al. 2010; Reshmi et al. 2017). Genetic susceptibility to Ph-like ALL has been linked to
some GATA3 polymorphisms, which are more frequent in individuals of US Hispanics and Native
180 JAK-STAT Signaling in Diseases

American genetic ancestry than in those of European descent (Perez-Andreu et al. 2013, 2015).
Large-scale comprehensive NGS studies have depicted the genomic landscape of Ph-like ALL. It is
now known that over 90% of Ph-like ALL harbor a diverse spectrum of genomic alterations
activating cytokine receptor genes and signaling kinase pathways, most of which converges to the
JAK-STAT signaling pathways in more than 70% of Ph-like ALL cases.

11.2.1.2 Down Syndrome-Associated ALL


Children with Down Syndrome (DS) carry a ~20-fold higher risk of developing ALL compared to
those without DS (Hasle et al. 2000). Infant ALL is extremely rare among the DS population and
the peak of incidence is slightly higher than in children without DS (Buitenkamp et al. 2014;
Dördelmann et al. 1998). Down syndrome-associated ALL (DS-ALL) carries a higher risk of
relapse compared to ALL patients without DS (Buitenkamp et al. 2014; Goto et al. 2011;
Whitlock et al. 2005; Zeller et al. 2005). In a retrospective analysis of 653 DS-ALL patients
enrolled in 16 international trials from 1995 to 2004, the 8-year cumulative incidence of relapse
was 26% ± 2%, compared to 15% ± 1% in those without DS (p < 0.001). The treatment-related
mortality is notoriously higher and mainly due to bacterial and viral infections (Buitenkamp et al.
2014; Shah et al. 2009; Whitlock et al. 2005; Zeller et al. 2005). Both these elements contribute to
the poor overall survival associated with DS-ALL (Buitenkamp et al. 2014; Chessells et al. 2001;
Whitlock et al. 2005). Precursor B-ALL is almost the exclusive ALL phenotype seen in DS-ALL
and is characterized by a distinct biology. The favorable cytogenetic features (such as ETV6­
RUNX1 translocation and high hyperdiplody) are less frequent and 40–45% of DS-ALL has
a normal karyotype (Buitenkamp et al. 2014; Forestier et al. 2008). As detailed below, DS-ALL is
associated with frequent alterations of the JAK-STAT signaling pathway.

11.2.1.3 CRLF2 Rearrangements and Concomitant JAK Mutations


Approximately half of Ph-like ALL and 60% of DS-ALL patients harbor CRLF2 rearrangements
(CRLF2r, Table 11.1) (Den Boer et al. 2009; Hertzberg et al. 2010; Mullighan et al. 2009b; Reshmi
et al. 2017; Roberts et al. 2014; Yoda et al. 2010). Two CRLF2 partners have been described in
these populations: P2RY8 and IGH, the resulting aberrant gene fusion leads to CRLF2 over-
expression (Den Boer et al. 2009; Mullighan et al. 2009b; Roberts et al. 2014). Rarely, P2RY8­
CRLF2 fusions can also be found in non-Ph-like ALL (Harvey et al. 2010; Reshmi et al. 2017).
A more rare CRLF2 alteration leading to CRLF2 overexpression is the CRLF2F232C activating
point mutation (Chapiro et al. 2010; Yoda et al. 2010). In about half of CRLF2r patients,
concomitant JAK1 or JAK2 mutations are found (Den Boer et al. 2009; Hertzberg et al. 2010;
Reshmi et al. 2017; Roberts et al. 2014; Yoda et al. 2010). They mostly affect the JAK2 Janus
homology domain (JH2) pseudokinase domain (which negatively regulates the JH1 kinase
domain), especially at the R683 residue, either as a point mutation (the most frequent being
JAK2R683G) or a 5-amino acid deletion removing R683 (JAK2ΔIREED) (Gaikwad et al. 2009;
Kearney et al. 2009; Malinge et al. 2007). JAK1 mutations are less common, the most frequent is
JAK1V658F, which is analogous to the JAK2V617F mutation found in MPN (Staerk et al. 2005).
Interestingly, both CRLF2r and JAK mutations are particularly prevalent in patients with DS-ALL
as they may occur in up to 30–50% of cases (Gaikwad et al. 2009; Hertzberg et al. 2010; Kearney
et al. 2009).
CRLF2 gene is located at the pseudoautosomal region 1 of Xp22/Yp11 and encodes for the
thymic stromal lymphopoeitin (TSLP) receptor, which is formed by the heterodimerization of
CRLF2 with interleukin-7 receptor-α (IL7R-α) on the cell membrane, and plays an important
role in early B-cell development (Scheeren et al. 2010). Upon its ligand binding, the physiolo­
gical downstream cascade involves STAT5 phosphorylation through JAK2 activation, leading to
dysregulated JAK-STAT signaling (Isaksen et al. 1999, 2002). In ALL patients’ cells, phospho­
JAK2 (pJAK2) and pSTAT5 are more elevated in the presence of CRLF2 alterations, regardless
JAK-STAT Signaling in Hematologic Malignancies 181

TABLE 11.1
JAK-STAT Alterations and Targeted Therapies in Acute Leukemias
Treatment options
Acute
leukemias JAK-STAT alterations Frequency Pre-clinical Clinical

Ph-like ALL
CRLF2 rearrangement 50% of Ph-like ALL, Type I and II JAK Ruxolitinib
(CRLF2r) 60% of DS-ALL inhibitors ± mTOR (NCT02723994)
± CRLF2F232C or MEK inhibitor. Ruxolitinib
± JAK1/2 alterations 50% of CRLF2r Givinostat, (NCT03117751 and
CART-cells NCT02420717
± IL7R-α alterations Around 5% of Ph-like
ALL
± SH2B3 alterations Individual cases
JAK2 fusion 5-10% of Ph-like ALL
EPOR rearrangement Around 5% of Ph-like Type I JAK
ALL inhibitors

DS-ALL
CRLF2r 60% of DS-ALL Type I and II JAK None
inhibitors ± mTOR
or MEK inhibitor.
Givinostat,
CART-cells
T-ALL
ETV6-JAK2 Individual cases Licochalcone A None
JAK1, JAK2, JAK3, Around 25% of T-ALL Ruxolitinib +
IL7R-α, PTPN2, FLT3, venetoclax (IL7R-α)
STATB5, SH2B3 and
IL2RB point mutations
AML
JAK mutations Individual cases of Ruxolitinib + Ruxolitinib (adults with
AMKL venetoclax (AML relapse AML regardless of
STAT3 activation Around 50% of adult regardless of JAK JAK alterations,
AML alterations) NCT00674479)

ALL: acute lymphoblastic leukemia, AML: acute myeloid leukemia, DS: Down Syndrome

of JAK mutation (Tasian et al. 2012). However, CRLF2 overexpression is probably not sufficient
on its own to drive leukemogenesis. In murine fetal liver cells, CRLF2 RNA-transduction is
associated with constitutive JAK2 and STAT5 phosphorylation as well as enhanced lymphoid
progenitor proliferation in the presence of growth factors (Russell et al. 2009). However,
CRLF2r-transduced Ba/F3 cells result neither in constitutive activation nor cytokine­
independent growth (Russell et al. 2009; Tasian et al. 2012). In addition, transgenic-mediated
CRLF2 overexpression in mice result neither in JAK-STAT pathway activation nor ALL
development (Kim et al. 2018).
Leukemogenesis is potentially driven by the cooperation between CRLF2r and JAK mutations,
as suggested by their co-occurrence. JAK mutations are usually not able to transform Ba/F3 cells
but can lead to cytokine-independent proliferation when co-expressed with erythropoietin (EPO)
or thrombopoietin (TPO) (Bercovich et al. 2008; Mullighan et al. 2009a). The co-expression of
P2RY8-CRLF2 and JAK2 mutations (JAK2R683G or JAK2P933R) in these cells result in constitu­
tive JAK2 and STAT5 activation as well as cytokine-independent proliferation (Hertzberg et al.
182 JAK-STAT Signaling in Diseases

2010; Mullighan et al. 2009; Yoda et al. 2010). In mouse models, transduction of CRLF2­
overexpressing/JAK2R683G or P933R fetal liver cells induce an aggressive leukemia with hepa­
tosplenic infiltration and circulating mutant progenitor B-cell blasts (Kim et al. 2018). The
same cooperation was also demonstrated between CRLF2F232C and JAK2 mutations (Bodegom
et al. 2012). JAK pharmacologic inhibition is effective in these preclinical models to abolish
hyperactive STAT signaling, cell proliferation, and in vivo tumor growth (Maude et al. 2012;
Mullighan et al. 2009b; Roberts et al. 2017; Tasian et al. 2012, 2017; Yoda et al. 2010).
Interestingly, in addition to JAK-STAT signaling, PI3K/mTOR pathway was also hyperacti­
vated in cells harboring CRLF2r and JAK mutation, providing another potential therapeutic
avenue (Tasian et al. 2012).
However, two studies challenge the driver effect of CRLF2r. In fact, P2RY8-CRLF2­
positive clone only represents a minor population at diagnosis and 30–50% of these patients
lose this clone at relapse (Morak et al. 2012; Vesely et al. 2017). Other genetic alterations
are persistent at relapse, particularly IKZF1 deletions, which become enriched in the re­
emerging clones. The above observations suggest that CRLF2r may be dispensable for
mediating relapse.

11.2.1.4 Other Mechanisms of JAK-STAT Activation in Ph-Like ALL


In a subset of Ph-like ALL patients without CRLF2r or JAK2 mutations, JAK2 fusions were
identified in 6% of children with NCI high-risk B-ALL, 7% of adolescents between 16 and
21 years old, 6% in young adults from 21 to 39 years old, and 8% in older adults from 40 to
86 years old (Jain et al. 2017; Pui et al. 2017; Roberts et al. 2014, 2016). At least 19 partner
genes have been reported, sharing the same mechanism consisting of in-frame fusions at the
N-terminus extremity of JAK2, which conserve the kinase ± the pseudokinase domains
(Kawamura et al. 2015; Reshmi et al. 2017; Roberts et al. 2012, 2014, 2016; Tirado et al.
2010; Yano et al. 2015). In vitro, the expression of JAK2 fusions results in constitutive STAT5
activation and cytokine-independent growth (Maude et al. 2012; Roberts et al. 2012, 2014;
Schinnerl et al. 2015).
In approximately 5% of patients with Ph-like ALL, rearrangements involving the EPO
receptor (EPOR) have been reported with four different partners: IGH, IGK, LAIR1, THADA
(Iacobucci et al. 2016; Reshmi et al. 2017; Roberts et al. 2012, 2014). The fusion of the first
three genes have been well described; these EPOR rearrangements result in the truncation of
EPOR’s C-terminus, which contains distal regulatory residues (Iacobucci et al. 2016). This
leads to stabilized expression of EPOR and hyperactive JAK-STAT, PI3K/mTOR, and
MAPK signaling. In mouse fetal liver cells, these rearrangements induce B-ALL development
(Iacobucci et al. 2016). Interestingly, analogous-truncated EPOR proteins are found in
autosomal benign erythrocytosis, which is caused by a germline mutation (de la Chapelle
et al. 1993).
IL7R-α gain-of-function is mediated by either amino acid 185 substitution or in-frame inser­
tion/deletion in less than 10% of Ph-like ALL (Roberts et al. 2014; Shochat et al. 2011). IL7R-α
mutations are often found in patients with CRLF2 fusions without concomitant JAK mutations.
Paired with CRLF2, the mutant IL7R-α leads to STAT5 hyperactivation and cytokine­
independent growth of progenitor lymphoid cells (Shochat et al. 2011).
Loss-of-function point mutations and small deletions of SH2B3 (Lnk) have been reported in
children with Ph-like ALL harboring CRLF2 alterations but without JAK or IL7R-α mutations
(Maude et al. 2012; Roberts et al. 2014). A recent report observed that SH2B3 deletions are more
enriched among B-ALL cases with intrachromosomal amplification of chromosome 21 (iAMP21)
(Baughn et al. 2018). Lnk protein, encoded by SH2B3, is a negative regulator of JAK-STAT;
therefore, SH2B3 deletion leads to hyperactive JAK-STAT signaling. In a mouse model, Lnk
deficiency synergizes with TP53 deletion to promote leukemogenesis via STAT5 activation (Cheng
et al. 2016). In xenograft models, IL7R-α and SH2B3-mutant ALL are sensitive to ruxolitinib
(Maude et al. 2012).
JAK-STAT Signaling in Hematologic Malignancies 183

11.2.2 JAK-STAT Pathway Alterations Are Commonly Found in Pediatric T-ALL


11.2.2.1 JAK2 Fusions
Although JAK2 fusions are more prevalent in Ph-like ALL, the former has been reported but
occurs in low frequency in T-lineage ALL. ETV6-JAK2 was initially described in a patient with
T-ALL and several other cases have since been published, which although are uncommon
(Lacronique et al. 1997). In T-ALL, the translocation fuses exon 5 of TEL and exon 19 of
JAK2, whereas in B-ALL, exon 4 of TEL is juxtaposed to exon 17 of JAK2 (Lacronique et al.
1997; Peeters et al. 1997). The aberrant transcript exhibits constitutive tyrosine kinase activity and
promotes STAT1, STAT5, PI3K, RAS/ERK, and NF-κB activation (Nguyen et al. 2001). In
several TEL-JAK2 transgenic mice models, the mice develop an invasive T-cell leukemia (Carron
et al. 2000; Santos et al. 2007; Schwaller et al. 1998).

11.2.2.2 JAK-STAT Pathway Mutations


Somatic mutations involving the JAK-STAT pathway are recurrent events in childhood T-ALL, but
only one cohort has investigated the prevalence of such alterations (Liu et al. 2017). In this study,
almost 25% of 264 patients harbor a point mutation involving this pathway, most commonly JAK3
(n = 20), IL7R (n = 18), PTPN2 (n = 13), FLT3 (n = 12), STATB5 (n = 11), JAK1 (n = 7), SH2B3
(n = 3), JAK2 (n = 3), and IL2RB (n = 1) mutations. Interestingly, 19 patients have several
concomitant mutations implicating this pathway, while 17 patients involve other pathways (Ras or
PI3K/AKT). The nature and functional impact of these alterations, particularly in the setting of
T-ALL, are largely unknown with a few exceptions. For example, 10 different JAK3 mutations have
been reported in childhood T-ALL, mainly missense, the most frequent being JAK3M511I. Most of
JAK3 mutations require both binding to the common γ-chain and the presence of JAK1 for their
transforming capacities (Degryse et al. 2014). In particular, JAK3M511I mutation alone may have
limited oncogenic potential in Ba/F3 cells as it is competing with wild-type JAK3 in one model.
Interestingly, a second JAK3 mutation was found in 30% of the adults studied and the double JAK3
mutants (JAK3M511I and JAK3A572T or JAK3A573V mutations) induce increased STAT5 phosphor­
ylation and cytokine-independent leukemia transformation. In the pediatric cohort cited earlier,
only 3 of the 30 patients contain a double JAK3 mutant but 50% harbors another potentially
oncogenic mutation (Liu et al. 2017). HOXA cluster rearrangements (leading to HOXA over-
expression) frequently co-occur with JAK3 mutations. In vivo, HOXA overexpression cooperates
with JAK3M511I to develop leukemia (Bock et al. 2014). Taken together, these data suggest that
JAK3 mutations are not “drivers” and require an additional event to promote leukemogenesis.
IL7R alterations are different to those seen in B-ALL. In T-ALL, they represent gain-of-function
indels within exon 6 (Zenatti et al. 2011). In most cases, these alterations introduce an unpaired
cysteine in the juxtamembrane/transmembrane extracellular region, which lead to dimerization of
IL7R mutants. In cell lines, they drive constitutive JAK1 signaling, cytokine-independent growth,
and promote viability; a phenotype which is reversed through type I JAK inhibitor and STAT5
inhibitor (Zenatti et al. 2011).
FLT3 mutations have been reported mainly in hyperdiploid pediatric T-ALL (Armstrong et al.
2004). These activating mutations were mostly present in the receptor-activation loop, a different
mechanism from the internal tandem duplication seen in AML (Armstrong et al. 2004). These FLT3
mutants were constitutively phosphorylated and transformed 32D cells (Yamamoto et al. 2001).
JAK1 mutations are missense substitutions located in the FERM domain (band 4.1, ezrin,
radixin, and moesin) or pseudokinase domain. In adults, the most frequent is JAK1V658F, which is
analogous to the JAK2V617F described earlier in B-ALL (Staerk et al. 2005). Among the seven
children reported within the largest cohort, this mutation was present in only one case and four
harbored the JAK1R724H mutation (Liu et al. 2017). Both mutations induced cytokine-independent
growth in Ba/F3 cell lines (Flex et al. 2008; Porcu et al. 2009), but these mutants need IL-9Rα or
IL-2Rβ expression for constitutive activation (Hornakova et al. 2009).
184 JAK-STAT Signaling in Diseases

The emergence of mutations can evolve under the selective pressure of therapy. In JAK3­
mutated murine cell lines, JAK1 mutations appear following treatment with type I JAK inhibitors
as a mechanism of resistance and, vice versa, JAK3 mutations emerge when JAK1-mutated cell
lines are treated with pan-JAK inhibitors (Springuel et al. 2014). This example suggests that
acquired JAK mutations may represent a mechanism of resistance to type I JAK inhibitors.

11.2.3 JAK-STAT Alterations Are Rarely Found in AML Despite Frequent STAT3
Phosphorylation
11.2.3.1 JAK Mutations
The current data suggest that the implication of the JAK-STAT pathway is less frequent in AML
than ALL. The genomic landscape of pediatric and adult AML did not reveal frequent recurrent
alterations of the JAK-STAT pathway genes (Bolouri et al. 2018; Ley et al. 2013). The canonical
MPN-associated JAK2V617F point mutation is uncommon in AML and may be present in around
5% of adult AML (Hidalgo-López et al. 2017; Jelinek et al. 2005; Steensma et al. 2005, 2006).
However, at least in some cases, it may be due to leukemic transformation of undiagnosed pre­
existing MPN.
Two JAK1 mutations (T478S, V623A) have been reported in adult but not in pediatric AML
patients (Tomasson et al. 2008). Although they lead to STAT1 phosphorylation, both are unable
to transform primary murine cells, suggesting that they act more as disease-modifier than as
driver mutations (Xiang et al. 2008).
Several JAK3 non-recurrent mutations have been reported in acute megakaryoblastic leukemia
(AMKL) occurring in patients with or without DS, including children (Hama et al. 2008; Kiyoi
et al. 2007; Malinge et al. 2008; Norton et al. 2007; Sato et al. 2008; Vita et al. 2007). Not all of
these mutations have been functionally characterized, but, in those that have been described, they
commonly lead to constitutive JAK3 activation, drive growth, and promote viability of Ba/F3 cell
lines (Malinge et al. 2008; Sato et al. 2008).

11.2.3.2 STAT3 Phosphorylation


Despite the paucity of JAK-STAT pathway gene alterations, constitutive STAT3 phosphoryla­
tion in adult AML blasts has been found in approximately 50% of patients and confers a poor
prognosis (Benekli et al. 2002; Xia et al. 1998). In children, the constitutive phosphorylation
of STAT3 and STAT5 on tyrosine 705 has been reported to be recurrent in a single study
(Redell et al. 2013). In this paper, STAT3 phosphorylation at basal state is not associated with
outcome, but a subgroup (11%) of patients, who displayed hyperphosphorylated STAT3 in
response to moderate doses of both granulocyte-colony stimulating factor and IL-6, has an
extremely favorable outcome (Redell et al. 2013). The pattern of STAT3 sensitivity changes at
relapse and represents a prognostic factor (Stevens et al. 2015). Furthermore, in pediatric
AML cell lines, pharmacologic inhibition of STAT3 led to apoptosis (Redell et al. 2011).
Collectively, these data highlight the important role of STAT3 activation in a subgroup of
pediatric AML even though the molecular mechanisms underpinning this dysregulation
remain unknown.

11.2.4 Therapeutic Implications of JAK-STAT Hyperactivation in Acute Leukemias


As JAK-STAT signaling pathways are constitutively dysregulated in a subset of Ph-like ALL
patients described above, abrogating its signal represents the logical therapeutic avenue. Several
approaches have been studied in preclinical models but only ruxolitinib has moved into clinical
trials for further investigation. The use of JAK inhibitors in T-ALL and AML is less explored but
will be briefly discussed.
JAK-STAT Signaling in Hematologic Malignancies 185

11.2.4.1 Type I JAK Inhibitor in Ph-like ALL


In vitro, type I JAK inhibitor ruxolitinib allows to limit, if not inhibit, cell proliferation
(Mullighan et al. 2009; Tasian et al. 2012; Yoda et al. 2010). Its efficacy has been also demon­
strated in Ph-like ALL patient-derived xenograft (PDX) models (Maude et al. 2012; Roberts et al.
2017; Tasian et al. 2017). Nevertheless, these preclinical studies also highlight potential challenges,
which may limit the benefits of ruxolitinib. First, ruxolitinib response in PDX models can be
variable depending on the underlying JAK-STAT pathway alterations, with JAK2 rearrangements
exhibiting a more robust sensitivity to ruxolitinib than CRLF2 rearrangements (Maude et al.
2012). Secondly, type I JAK inhibitor may induce paradoxical JAK2 hyperphosphorylation
conferring resistance to ruxolitinib (Weigert et al. 2012). Thirdly, in genetically engineered mouse
models, after a transient response to JAK2 inhibition, mutant B-cells remain viable by upregulat­
ing c-Myc expression (Kim et al. 2018). However, the combination of BET bromodomain
inhibitors with ruxolitinib can abrogate this effect (Kim et al. 2018). Lately, JAK-STAT hyper-
signaling may be toxic to cells and some patients have a therapeutic-emergent loss-of-function of
USP9X, an adaptive mechanism to limit this effect (Schwartzman et al. 2017). In this study,
tempering JAK-STAT signaling paradoxically improved proliferation; hence, if present in vivo, this
effect could limit the efficacy of suboptimal JAK inhibition. The clinical relevance of all these
phenomena remains to be demonstrated and ongoing clinical studies will address the efficacy of
ruxolitinib in this population.
Three clinical trials are currently investigating the safety and efficacy of the combination of
ruxolitinib and conventional chemotherapy backbone. The Children’s Oncology Group (COG)
phase 2 trial AALL1521 (NCT02723994) is active and enrolls NCI high-risk Ph-like ALL patients
with JAK-STAT pathway lesions who are stratified according to the underlying genetic alterations
(CRLF2r, JAK mutations, JAK2 fusions or EPOR rearrangements) and end-induction MRD
status. The first phase of this trial has determined that continuous dosing of ruxolitinib at 40mg/
m2/dose for 28 days per cycle in combination with intensive multi-agent chemotherapy regimen is
safe and tolerable in children and young adults with eligible Ph-like genetic lesions (Tasian et al.
2018). St. Jude Children’s Research Hospital’s Total therapy XVII protocol (NCT03117751) adds
ruxolitinib to chemotherapy in children with JAK–mutant Ph-like ALL. For adults, a phase 2
MD Anderson Cancer Center trial (NCT02420717) is evaluating ruxolitinib for relapsed/refrac­
tory JAK–mutant ALL. Patients are initially treated with up to 3 weeks of ruxolitinib as
monotherapy and multiagent chemotherapy will be eventually added for those with incomplete
response.

11.2.4.2 Others Approaches in Ph-Like ALL


Type II JAK inhibitors may represent an alternative to allow more profound inhibition of JAK­
STAT signaling. The compound CHZ868 was able to abrogate JAK-STAT signaling and prolonged
JAK–mutant PDX mice survival. However, similar to ruxolitinib, the emergence of resistance
mediated by acquired point mutations was also documented (Wu et al. 2015). The combination of
JAK inhibitors with HSP90 or BET bromodomain inhibitors has been tested in preclinical models
and has shown to be effective in inhibiting JAK-STAT signaling in Ph-like ALL. In addition to
JAK-STAT pathway, PI3K/mTOR and Ras-MAPK pathways are also activated in Ph-like ALL cell
lines with CRLF2r and JAK mutations, the addition of a pharmacological inhibition of the mTOR
pathway to ruxolitinib was investigated. It resulted in improved inhibition of leukemia growth in
Ph-like ALL PDX models (Tasian et al. 2017). Furthermore, the addition of selumitinib, a MEK
inhibitor, to AZD1480, a JAK1/2 inhibitor, markedly improved the inhibition in cell lines but not in
PDX mice (Suryani et al. 2015). The histone deacetylase inhibitor givinostat (ITF2357) has shown
potent efficacy as monotherapy to treat CRLF2r Ph-like ALL in vitro and in PDX mice (Savino
et al. 2017). Finally, chimeric antigen receptor T-cells targeting TSLP receptor has demonstrated
efficacy to eradicate CRLF2-overexpressing ALL in PDX mice without off-target effect (Qin et al.
2015).
186 JAK-STAT Signaling in Diseases

11.2.4.3 JAK Inhibitors in T-ALL and AML


In a preclinical study, the combination of ruxolitinib to the BCL2 inhibitor venetoclax has shown
to control cell proliferation of IL7R-α-mutated T-ALL in vivo and in vitro, in contrast to
ruxolitinib alone. This therapeutic combination has not yet been investigated in patients (Senke­
vitch et al. 2018). Another study has found the same combination to be efficacious in AML,
regardless of the underlying genomic alteration, which could be explained by effective targeting of
the AML niche (Karjalainen et al. 2017).
A phase II study (NCT00674479) investigated ruxolitinib as a monotherapy for adult patients
with relapsed/refractory acute leukemias, including 28 with AML (which was secondary to MPN
in 10 cases), irrespective of JAK-STAT mutation status. Despite being well tolerated, the drug has
only modest effect; among three patients (all with secondary AML) who achieved a complete
response, two had JAK2V617F mutation (Eghtedar et al. 2012). A phase I/II in adult AML found
similar results and no data are available for pediatric AML (Pemmaraju et al. 2015).

11.2.5 Conclusion
JAK-STAT signaling is involved in a significant proportion of the three leukemia types presented here
although the mechanisms differ between each disease entity. Moreover, the clinical impact of JAK­
STAT involvement may vary between these diseases. In AML, genetic alterations involving this
pathway are rare but functional dysregulation is frequently associated. The underlying mechanism
remains largely unknown and if this element represents the driver event or only secondary to other
alterations is yet to be determined. Nevertheless, the implication of JAK-STAT pathway has been
associated with prognosis in AML, highlighting its contribution to leukemogenesis.
In T-ALL, JAK-STAT alterations comprise a quarter of cases through different mechanisms
and genes. Interestingly, these alterations are frequently associated with each other as well as with
other pathways. This illustrates the cooperation between different oncogenes to foster cell
proliferation or survival.
Ph-like ALL is the most characterized and studied disease entity, harboring frequent and
heterogeneous JAK-STAT pathway alterations. Its discovery presents a unique opportunity for
molecularly targeted therapy to improve the unfavorable outcomes of this patient population. The
therapeutic incorporation of JAK inhibitors in different preclinical studies has shown promising
responses despite several challenges have since arisen. The ongoing clinical studies will allow
confirming the safety and efficacy of ruxolitinib as a clinically active JAK inhibitor in Ph-like ALL.

11.3 Myeloproliferative Neoplasms


Myeloproliferative neoplasms (MPN) are clonal hematopoietic stem cell disorders resulting in the
excessive production of cytokine-independent mature myeloid blood cells (Tefferi 2006). Under
the WHO 2016-revised classification (Arber et al. 2016), they are divided into four distinct
“classic” MPN: chronic myelogenous leukemia (CML), polycythemia vera (PV), essential throm­
bocythemia (ET), and primary myelofibrosis (PMF) (Swerdlow et al. 2008). CML is a distinct
entity within the MPN spectrum; since BCR-ABL fusion has no involvement with the JAK-STAT
pathway, it will thereby not be discussed. Three other entities belong to MPN in this classifica­
tion: chronic eosinophilic leukemia not otherwise specified, chronic neutrophilic leukemia, and
unclassifiable MPN, which are all exceptionally rare in childhood. This chapter will focus on PV,
ET, and PMF in children.
Although being the more prevalent MPN in this age group, PV and ET remain extremely rare
in children. The annual incidences of PV and ET are estimated around 0.2 and between 0.1 and
0.4 per million of person under 20 years old, respectively (Hasle 2000; Hasle et al. 1995, 1999;
Jensen et al. 2000; McNally et al. 1997), which correspond to an incidence of 40–90-fold lower
JAK-STAT Signaling in Hematologic Malignancies 187

than in adults (Teofili et al. 2008). There is a female predominance in both adults and children
with PV (Chim et al. 2005; Giona et al. 2012). PMF is uncommon in children with less than 50
cases published in the literature to date (DeLario et al. 2010, 2012). With regards to the clinical
picture, the disease presentation in children is distinct from adult patients. PV is symptomatic in
the majority of adult patients and thrombotic events occur frequently (Marchioli et al. 2005;
Tefferi et al. 2013), whereas, in children, less than half will present with clinical symptoms and
thrombosis is rare (Giona et al. 2012). The prognosis also differs according to age and death is
estimated in around 8% of adults but none in children (Giona et al. 2012; Marchioli et al. 2005;
Tefferi et al. 2013). ET is diagnosed mainly as an incidental finding on blood counts in children
(Giona et al. 2012), but in adults, 50–70% of patients present with symptoms at diagnosis such as
asthenia, vasomotor symptoms, bleeding, and thrombosis (Cervantes et al. 2002; Chim et al.
2005; Cortelazzo et al. 1990; Fenaux et al. 1990). PMF occurs mainly in toddlers with frequent
transfusion support. The progression to leukemia is a major concern in adults with PMF but has
never been reported in children (Saksena et al. 2014). Spontaneous resolution of the disease in
children has been described in single case reports but the only published cohort reports 26% of
spontaneous remission among 19 patients (DeLario et al. 2012). Due to these important
differences in the disease evolution, data from adult studies should be carefully interpreted and
should not be extrapolated to children.
Cases of polycythemia, thrombocytosis, and myelofibrosis occurring in childhood comprise
a higher proportion of hereditary disorders due to a constitutional defect. Notably, germline
mutations in MPL, THPO, VHL, PHD2, and HIF-2α have been reported in patients with
hereditary polycythemia or thrombocytosis (Ang et al. 2002; Ding et al. 2004; Lenglet et al.
2018; Semenza 2009; Skoda 2010; Teofili et al. 2010). These rare monogenic defects lead to
polyclonal proliferation with a high penetrance and early disease onset in childhood. As they
do not meet the classical definition of MPN, they will not be discussed here. These conditions
are distinct from the “hereditary” MPN, but the boundary between these two entities are
becoming more porous as clonal hematopoiesis may also be found in hereditary polycythemia
(Kralovics et al. 2003); but for didactical purpose, the distinction, even if somewhat artificial,
will be kept here.

11.3.1 MPN-Associated JAK2 Mutations Are Less Frequent in Children than in Adults
11.3.1.1 Epidemiology of JAK2V617F Mutation
The first molecular defect identified in PV was the loss-of-heterozygosity (LOH) at 9p in
approximately a third of adults (Kralovics et al. 2002). The minimal amplified region contained
several genes, including JAK2, of which the specific implication was initially not demonstrated.
A recurrent somatic guanine-to-thymidine substitution of nucleotide 1849 within the exon 14 of
JAK2 was simultaneously identified in 2005 by several teams. This missense heterozygous
mutation results in the substitution of valine to phenylalanine at codon 617 (JAK2V617F) and
was found in more than 85% of adults with PV and 50% of those with ET and PMF (Baxter et al.
2005; James et al. 2005; Kralovics et al. 2005; Levine et al. 2005; Lippert et al. 2006). It
preferentially occurs in a common germline 46/1 (also called GGCC) JAK2 haplotype, present
in about half of the population (Jones et al. 2010; Pietra et al. 2012).
In children, due to the smaller cohort studied, the frequency of this mutation is less precisely
known but, as in adults, it frequency varies among the different MPN (Table 11.2). In PV, four
studies identified JAK2V617F to be present in 25–37% of children with sporadic PV (maximum
population studied: 11 patients), while another study reported a frequency as high as 75% among
eight cases (Cario et al. 2008; Giona et al. 2012; Ismael et al. 2012; Teofili et al. 2007a) Interestingly,
this mutation can be detected in congenital PV but not in familial PV (Kelly et al. 2008). This fact
highlights the difficulty to assess the real etiology of numerous cases of polycythemia with
subnormal EPO levels and a significant proportion of these cases may not represent a real MPN.
In fact, not all the cases studied (with the exception of the last paper) had documented evidence of
188 JAK-STAT Signaling in Diseases

TABLE 11.2
JAK-STAT Alterations and Targeted Therapies in Myeloproliferative Neoplasms
Treatment options FDA-
Myeloproliferative JAK-STAT approved
neoplasms alterations Frequency in children Pre-clinical Clinical therapy

Polycythemia vera
JAK2V617F 25-75% of cases Type I JAK Ruxolitinib in adults, Ruxolitinib
mutation inhibitors 2 children reported
Exon 12 JAK2 Individual cases
mutations (n=2)
SH2B3 mutations 2 cases among 320
MPN
Essential
thrombocythemia
JAK2V617F 22-56% of cases Type I JAK Ruxolitinib in adults,
mutation inhibitors no experience in
MPL mutations Individual cases children
(n=2)
CALR mutations 23.5% of cases
(n=8/34)
SH2B3 mutations 6 cases among 320
MPN
Primary
myelofibrosis
JAK2V617F Not reported Type I JAK Ruxolitinib and other Ruxolitinib
mutation (n=0/17) inhibitors type I JAK inhibitors
CALR mutations 50% of cases (pacritinib, memeloti­
(n=7/14) nib and fedratinib) in
SH2B3 mutations 15 cases among 320 adults, no experience in
children
MPN
MPN: Myeloproliferative neoplasms

clonal disease with clonal hematopoiesis and/or endogenous erythroid colonies and/or PRV-1
RNA overexpression (Cario et al. 2008; Teofili et al. 2007b). Despite these, the cases fulfilled
the WHO 2016-revised criteria for PV (Arber et al. 2016). Due to important differences
between pediatric and adult PV, these criteria may not apply accurately in children than in
adults. It has been suggested that, in the absence of JAK2V617F mutation in a child with
sporadic PV, the initial WHO 2001 criteria (made before the identification of JAK2V617F and
incorporating other indirect evidence of MPN) may be more appropriate to accurately make
the diagnosis of PV (Barbui et al. 2014; Tefferi et al. 2007; Teofili et al. 2008; Vardiman et al.
2002).
In pediatric ET, the prevalence of JAK2V617F mutation is less different than that of adults,
from 22% to 56% gathered from the eight published studies, the largest enrolling 63 patients (Fu
et al. 2016; Giona et al. 2012; Ismael et al. 2012; Randi et al. 2015; Sekiya et al. 2016; Teofili
et al. 2007a, 200, 2009). This mutation has not been reported in hereditary ET but, interestingly,
a dominant germline JAK2V617I mutation was found in a family with hereditary thrombocytosis
(Mead et al. 2012). Sporadic ET represents a similar challenge than PV with regards to the
diagnostic criteria. Some patients with persistent thrombocytosis may have other underlying
mechanisms than clonal proliferation, which point to the MPN (Fu et al. 2013; Randi et al.
2015). As with PV, it should be emphasized to rule out a hereditary cause of thrombocythemia
(Teofili et al. 2008).
JAK-STAT Signaling in Hematologic Malignancies 189

In summary, a significant proportion of cases of polycythemia and thrombocythemia in child­


hood may not be related to MPN despite biological findings suggest no other causes, which render
the interpretation of cohort and biological studies more difficult. Among cases of polycythemia and
thrombocythemia with markers of MPN, the JAK2V617F mutation is not as frequent as in adults.
Due to the extreme rarity of PMF among children, very few data are available. A single study
found no JAK2V617F mutation among 17 children (DeLario et al. 2012).

11.3.1.2 JAK2V617F Effect on Downstream Signaling


JAK2V617F is located in the JH2 domain (Ungureanu et al. 2011). The point mutation seems to
destabilize the JH1-JH2 interaction, and acts as a gain-of-function mutation resulting in consti­
tutive activation of not only JAK2 and STAT, but also of MAPK and PI3K pathways (James
et al. 2005; Silvennoinen and Hubbard 2015; Wernig et al. 2006; Zhao et al. 2005). Transduced in
cell lines, JAK2V617F promotes cytokine-independent growth and excess proliferation in the
presence of EPO, IL-3, and IGF-1 (James et al. 2005; Levine et al. 2005; Staerk et al. 2005;
Zhao et al. 2005). In murine models, this mutation induces a myeloproliferative PV-like disorder
manifested by splenomegaly, bone marrow hyperplasia, and EPO-independent erythrocytosis
within the first 3 months (James et al. 2005; Lacout et al. 2006; Wernig et al. 2006). Subsequently,
an important myelofibrosis develops, causing further increase in splenomegaly, followed by
regression of polycythemia and development of cytopenias (Lacout et al. 2006).
The clinical phenotype of MPN probably depends on the intensity of the JAK-STAT signaling,
linked to the number of JAK2V617F copies. Transgenic mice with a higher JAK-STAT signal
intensity develop a PV phenocopy, whereas lower intensity results in an ET-like disease with
mainly thrombocytosis (Tiedt et al. 2008). These in vivo data support the observation that most
patients with PV harbor a 9p LOH leading to homozygous JAK2V617F in contrast to the
heterozygous mutation in patients with ET (Dupont et al. 2007; Scott et al. 2006).
In addition, the mechanism contributing to phenotype variation may be related to the activation of
a specific downstream effector. PV, ET, and PMF are associated with preferential phosphorylation of
STAT5, STAT1, and STAT3, respectively (Chen et al. 2010; Teofili et al. 2007c). Consistently, in
a murine model, STAT1 knockout results in decreased thrombocytosis and favors erythropoiesis
(Duek et al. 2014). The fact that the same nucleotide substitution is found in three different diseases
with different downstream effects is currently not completely understood. Other modifying factors,
such as host polymorphisms or other cooperating mutations, may contribute to the phenotype
diversity and are currently under investigation.
Despite these phenotype variations, STAT5 probably plays a central role in all MPN. pSTAT5
is elevated in each phenotype and even in the absence of JAK2V617 mutation (Teofili et al. 2007c).
In vitro, both STAT5 and PI3K activation are required for the EPO-independent growth induced
by JAK2V617F (Garçon et al. 2006; Ugo et al. 2004). In mice models, STAT5 activation is required
for the development of JAK2V617F-induced MPN (Walz et al. 2012; Yan et al. 2012). The high
pSTAT5 in PV is partially linked to its capacity to upregulate and cooperate with the anti­
apoptotic protein Bcl-xL, which is also upregulated in PV (Garçon et al. 2006; Silva et al. 1998).
PI3K/Akt/mTOR pathway is activated downstream of JAK2 activation. In JAK2V617F cell lines,
the pharmacologic inhibition of AKT hampered cell growth (Khan et al. 2013). The combination
of PI3K and JAK1/JAK2 inhibition reduce spleen volume in mice inoculated with JAK2V617F Ba/
F3 cells (Bartalucci et al. 2013; Khan et al. 2013).
Whether JAK2V617F is sufficient by itself to initiate MPN is controversial. This mutation has
been identified in primitive CD34+ CD38− HSC and skews the maturation toward erythroid
differentiation (Jamieson et al. 2006). However, a study found no expansion of this compartment
in the sole presence of JAK2V617F mutation (Anand et al. 2011). In a mouse model, the presence
of the thrombopoietin receptor (MPL) (but not thrombopoietin) was mandatory to induce the
MPN phenotype in a JAK2V617F background (Sangkhae et al. 2014). This highlights the fact that
this mutation may not be a driver and requires additional cooperative event(s). Furthermore,
190 JAK-STAT Signaling in Diseases

MPN clonality may not correlate with JAK2V617F allele burden (Nussenzveig et al. 2007). A study
of families with more than one person affected by MPN showed that in some families,
a relative may have a JAK2V617F-positive MPN, while another family member is afflicted
with a JAK2V617F-negative MPN (Bellanné-Chantelot et al. 2006). In the same line, JAK2V617F
allele may be detected at a low frequency in the general population without evidence of MPN
(Martinaud et al. 2010; Nielsen et al. 2011; Sidon et al. 2006; Xu et al. 2007). In conclusion,
the contribution of this mutation to the pathogenesis of MPN has been demonstrated in
several in vitro and in vivo models. Nevertheless, this molecular defect probably does not
explain entirely the pathophysiology of MPN. Further mechanistic studies are still needed to
better define its functional role.

11.3.1.3 Exon 12 JAK2 Mutations


Among the subset of patients with JAK2V617F-negative PV, some present with mutations in the
exon 12 of JAK2 (Tefferi 2010). Initially described in 2007 (Scott et al. 2007), more than 15
heterozygous gain-of-function mutations have been reported to date in adults, including point
mutations, in-frame deletions, and duplications, mainly affecting seven highly conserved amino
acid residues (F537–E543) within exon 12 (Butcher et al. 2008; Pardanani et al. 2007; Passamonti
et al. 2011; Pietra et al. 2008). These mutation carriers tend to be adult patients with a familial
history of PV and mostly present at a younger age with an isolated erythrocytosis, a subnormal
serum erythropoietin level, and a similar prognosis to that of JAK2-mutant PV (Martínez-Avilés
et al. 2007; Pardanani et al. 2007, 12; Passamonti et al. 2011). To date, only two children have
been reported with an exon 12 JAK2 mutation (Cario et al. 2008). In Ba/F3 cell lines, among the
four mutations tested, all resulted in IL-3-independent growth (Scott et al. 2007). In a retroviral
transfer murine model, these mutations led to a myeloproliferative disease with marked erythro­
cytosis and modest thrombocytosis (Scott et al. 2007).

11.3.2 Other Genomic Alterations Activating JAK-STAT in PMF and ET


11.3.2.1 MPL Mutations
Among adults with ET and PMF, less than 10% harbor a MPL mutation with a slightly different
clinical phenotype: older age, higher platelet counts but lower hemoglobin level, and microvascular
symptoms associated with arterial thrombosis over the course of disease, and similar survival rates
(Beer et al. 2008; Pardanani et al. 2006; Pikman et al. 2006; Vannucchi et al. 2008). Several somatic
MPL mutations have been reported since 2006 in ET and PMF, the most frequent being MPLW515L,
which has been reported in two children with ET (Beer et al. 2008; Farruggia et al. 2013; Pardanani
et al. 2006; Putti et al. 2017; Teofili et al. 2010). These patients display functional similarities with
patients carrying a germline MPLS505N mutation but follow a different disease course; the latter may
evolve towards myelofibrosis but no leukemic transformation has been reported and hematopoiesis
remains polyclonal (DeLario et al. 2012; Ding et al. 2004; Teofili et al. 2007a, 2010).
MPL encodes for the thrombopoietin receptor and plays a key role in megakaryocyte prolifera­
tion and survival. Both somatic and germline mutations are located in the exon 10 and result in
a gain-of-function of the protein. The mutations are mainly heterozygous but acquire homozygosity
through uniparental isodisomy as previously described (Szpurka et al. 2009). In case of a somatic
mutation, the origin of clonal mutations was set at the multipotent HSC level (Chaligné et al. 2007).
Exon 10 is located in the juxtamembrane domain of MPL, which has an auto-inhibitory role,
limiting MPL activation (Staerk et al. 2006). Activating mutations result in loss of the inhibitory
signal and constitutive activation, despite the absence of thrombopoietin binding (Staerk et al.
2006). In several cell lines (Ba/F3, 32D and UT7), transduction of MPLW515L, but not of all
mutations found in patients, led to cytokine-independent growth (Chaligné et al. 2008; Pikman
et al. 2006). In Ba/F3 cell lines, MPLS505N results in cytokine-independent proliferation (Ding et al.
2004). As in normal physiologic states, MPL dimerization induced by TPO leads to JAK2
JAK-STAT Signaling in Hematologic Malignancies 191

phosphorylation (Drachman et al. 1995; Tortolani et al. 1995), in MPLW515L Ba/F3 cells, JAK­
STAT pathway was activated not only with JAK2, STAT3, STAT5, but also AKT and ERK being
phosphorylated (Pikman et al. 2006). Moreover, MPLS505N Ba/F3 cells result in constitutive MEK1/2
and STAT5B phosphorylation in the absence of IL-3 (Ding et al. 2004). In a mouse model,
transplantation of MPLW515L bone marrow leads to a MPN-like phenotype characterized by
splenomegaly, splenic infarct, marked thrombocytosis, and reduced life expectancy (Pikman et al.
2006). The contribution in the pathogenesis of these somatic mutations is yet not known in patients.
Some cases harbor several MPL mutations or concomitant JAK2V617F within the same clone (Lasho
et al. 2006). The clinical impact of these two mutations differs as patients with germline MPLS505N
may evolve toward myelofibrosis without leukemic transformation and hematopoiesis remains
polyclonal (Teofili et al. 2010).

11.3.2.2 CALR Mutations


CALR gene encodes calreticulin, a Ca2+ binding protein associated with multiple functions, including
proper protein folding in the endoplasmic reticulum (ER) and calcium homeostasis (Eggleton and
Michalak 2013). Through whole exome sequencing, somatic CALR mutations have been identified in
2006 and may be present in up to 70% of adults with ET or PMF without JAK2 or MPL mutations,
but not in patients with PV (Klampfl et al. 2013; Nangalia et al. 2013). In children, they were found in
23.5% among 34 ET cases and 50% in 14 PMF cases (An et al. 2014; Giona et al. 2014). Despite the
rarity of these mutations, they still occur more frequently than MPL mutations in ET and PMF for
both children and adults (Tefferi et al. 2014a). More than 50 mutations have been described; they
consist of insertions and deletions located in exon 9 of CALR (Klampfl et al. 2013; Nangalia et al.
2013). The most frequent CALR mutations in adults are the 52-base pair (bp) deletion (type 1) and
the 5-bp insertion (type 2 mutation), encompassing more than 80% of CALR mutations (Rumi et al.
2014a, 2014b). In children with PMF, only type 2 mutations were reported whereas both types were
described in pediatric ET (An et al. 2014; Giona et al. 2014).
The pathophysiological relationship between CALR, a housekeeping chaperone protein, and
MPN was not obvious. Studies have demonstrated that mutant CALR resulted in abnormal
physical interaction with the MPL protein inside the ER (Araki et al. 2016; Chachoua et al.
2016). The coupled proteins are exported on the cell surface and result in constitutive TPO
activation leading to JAK2 dimerization and activation (Araki et al. 2016; Chachoua et al. 2016).
In a mouse model, type 1 and type 2 mutants induce thrombocytosis with frequent progression to
myelofibrosis (Marty et al. 2016). This phenotype was more frequently and rapidly induced by
type 1 than type 2 mutations. MPL, but not TPO, is mandatory in the disease progression of these
mice (Marty et al. 2016). In adults with ET, CALR mutations (especially type 2) are associated
with higher platelet counts, lower hemoglobin and leukocyte counts than those with JAK2V617F,
suggesting that differences in JAK-STAT dysregulation depends on the involved gene and the type
of CALR mutation (Tefferi et al. 2014b). Coexistence of CALR and JAK2V617F mutations is rare
but has been reported (Lundberg et al. 2014).

11.3.2.3 SH2B3 Mutations


Five to seven percent of MPN patients harbor SH2B3 mutations (McMullin and Cario 2016).
The SH2B3 gene is located on chromosome 12q24.12, encodes for the protein LNK, which
belongs to the LNK/SH2B family of adaptor proteins, which plays an important role in early
hematopoiesis. Most of SH2B3 mutations are missense mutations that predominantly target exon
2 coding for the Pleckstrin homology (PH) domain, resulting in a reduced level of protein activity.
The other reported SH2B3 alterations consist of deletions of exons 2 and 5, targeting once again
the PH domain (Maslah et al. 2017). Functional analyses demonstrate that these SH2B3 mutants
in vitro exhibit aberrant JAK2/STAT3/STAT5 activation as a result of decreased LNK activity
(Oh et al. 2010). Therefore, LNK acts a negative regulator of the JAK-STAT pathway. SH2B3
192 JAK-STAT Signaling in Diseases

mutations can be found in all MPN subtypes and can co-exist with other MPN driver genes
(JAK2, MPL, CALR) or other genes (TET2, CBL) (Maslah et al. 2017). Interestingly, it has also
been observed that MPN patients may develop SH2B3 mutations during leukemic transforma­
tion, during which their frequency increases to 13% (Maslah et al. 2017; Pardanani et al. 2010).
This suggests that SH2B3 mutations may contribute to the leukemic transformation of MPN.
However, the method used in these studies has low sensitivity and may not detect subclonal
SH2B3 mutations during the chronic phase of MPN.

11.3.2.4 Mutations in Other Genes and Interaction between Co-existing Mutations


Adult patients without JAK2, MPL and CALR mutations represent less than 10% of PMF and
around 15% of ET; they are often referred as triple-negative MPN. Alterations in other genes have
been described but without functional impact on the JAK-STAT pathway (Cabagnols et al. 2016;
Feenstra et al. 2016). For example, mutations in TET2 have been reported in both JAK2V617F­
positive and JAK2V617F-negative patients, in addition to mutations in CBL, ASXL1, IDH, IKZF1
and EZH2 in triple-negative patients (Guglielmelli et al. 2014; Tefferi and Vainchenker 2011).
They represent a small subset of adults and their functional consequences remain under investiga­
tion. In children, the potential implication of these genes stays unknown.
Lastly, the mutations described herein appear at various stages of clonal evolution and may co­
exist with each other (Lundberg et al. 2014). This has a potential impact on the disease course
and the risk of leukemic transformation in adults (Klampfl et al. 2011; Lundberg et al. 2014). The
impact on the JAK-STAT pathway for each specific alteration or combination is not entirely
known. However, genomic analyses showed that for all adults with MPN, regardless of the
underlying mutation, a transcriptional signature consistent with activated JAK2 signaling is
present (Rampal et al. 2014).

11.3.3 Therapeutic Opportunities of JAK-STAT Inhibition in MPN


MPN in adults was the first field leading to the development of type 1 JAK inhibitors. Currently,
ruxolitinib is the only drug with the Food and Drug Administration (FDA) and European
Medicine Agency (EMA) approval for PMF as first-line therapy since 2011 and then hydroxyurea
(HU)-resistant or -intolerant PV since 2014.
In HU-resistant or -intolerant PV in adults, ruxolitinib has shown to be superior to placebo in
improving symptoms, splenomegaly and hematocrit (Passamonti et al. 2017; Vannucchi et al.
2015). Data in children is limited to a 26-month old patient with a JAK2V617F PV-associated
Budd-Chiari syndrome and another isolated patient enrolled on the COG ADVL1011 phase 1
trial (NCT01164163) (Coskun et al. 2017; Loh et al. 2015).
In adults with HU-resistant or -intolerant ET, ruxolitinib allows clinical and biological
improvement but the impact on hemorrhage, thrombosis and leukemic transformation is yet
unclear (Gunawan et al. 2018; Harrison et al. 2017a; Verstovsek et al. 2017). No experience has
been published in children with HU-resistant or –intolerant ET.
Ruxolitinib has shown efficacy in adults with PMF, allowing to reduce clinical symptoms and
spleen volume (Al-Ali et al. 2016; Verstovsek et al. 2010). As previously evocated, PMF is rare
and has a markedly more benign evolution in the pediatric population. Therefore, no pediatric
experience of ruxolitinib in PMF is yet published.
More recently, other type I JAK inhibitors have demonstrated significant clinical benefits in adult
PMF (such as pacritinib, memelotinib and fedratinib) (Harrison et al. 2017b; Mascarenhas et al.
2018; Mesa et al. 2017; Pardanani et al. 2015), but there is no such experience for these therapies in
children.
In conclusion, JAK inhibitors have become the standard treatment for adult MPN, as first
or second-line therapies. However, it is not clear whether these therapies are truly disease-modifying
and improve outcomes (Cervantes and Pereira 2017). The symptoms’ improvement may also be
JAK-STAT Signaling in Hematologic Malignancies 193

linked to the anti-inflammatory effect of ruxolitinib (Hasselbalch 2013). Moreover, MPN clones
persist during therapy and adaptive mechanisms of resistance through paradoxical JAK-STAT
activation have been described (Koppikar et al. 2012). To overcome these pitfalls, others strategies
including combinatorial therapy between ruxolitinib and novel agents (eg. histone deacetylase
inhibitors) are being developed for adult patients but their benefits have not been established in
children (Bose and Verstovsek 2017).

11.3.4 Conclusion
The JAK-STAT pathway plays an important role in the pathogenesis of the three classical MPN
discussed above. Functional analyses showed the central role of this pathway in the disease evolution.
The identification of the JAK2V617F mutation was an important breakthrough deepening our under­
standing of the underlying causes of the disease. However, the molecular mechanisms and the primum
movens remain not known precisely. Other genetic factors seem to contribute to the disease onset and
evolution. In children, the knowledge of these studies pertains only to a subgroup of patients and
a significant proportion of children do not have a molecular or mechanistic explanation of their
MPN. Investigations are needed in this specific population to address this question. Finally, the
paucity of MPN in children, the distinct evolution and the diverse genomic spectrum have halted the
development of targeted therapies in this population. To date, the experience is limited; international,
multicentric collaborations will be needed to allow a comprehensive pediatric-tailored management
for MPN.

11.4 Lymphomas
The lymphoproliferative neoplasms encompass a large category of malignant lymphoid prolifera­
tions arising from B- or T-lymphocytes at various maturation stages. The 2016 WHO revised
classification distinguishes more than 40 entities. Lymphomas are classically divided into two
broad categories: Hodgkin lymphomas (HL) and non-Hodgkin lymphomas (NHL).
HL are the most frequent lymphomas among children with a peak incidence during adolescence
(30.1 cases per million between 15 and 19 years in the USA) (Noone et al. 2018). The most
frequent histologic subtype is classical HL (cHL). Longterm survival rates in pediatric HL now
reaches above 90% with contemporary treatment regimens but are negatively affected by treat­
ment-related toxicities, especially in high-risk disease being exposed to intensified combinatorial
chemo-radiotherapy (Smith et al. 2010; Thomson and Wallace 2002). The dysregulated pathways
of lymphomatous cells are partly derived from genetic influences and partly as the result of the
important interaction with the microenvironment surrounding the tumoral cells (Mathas, Hart­
mann, and Küppers 2016). The impact of age on the biology of HL is poorly understood. Clinical
observations often indicate that older children present with more extensive disease, despite the
comparable survival rates among all age subgroups (Englund et al. 2018). However, this difference
may be related to the difference in growth and development rather than the underlying disease
biology. Thus, the majority of biological studies in HL are not age-based and the discussed
literature herein includes representation from the adult population including adolescents and
young adults.
NHL are less frequent in childhood than adulthood as the incidence rises with increasing age,
from 7.4 per million between 1 and 4 years to 20 per million between 15 and 19 years in the USA
(Noone et al. 2018). Pediatric NHL is markedly different from their adult counterpart, being
almost exclusively high-grade lymphomas and mostly of B-cell origin (Burkhardt et al. 2005).
Almost all NHL subtypes are less common during childhood with the exception of sporadic
Burkitt lymphoma (BL). No JAK-STAT pathway implication has been demonstrated in BL;
hence, this entity will not be discussed here. Diffuse large B-cell lymphoma (DLBCL) is
the second most frequent NHL in childhood. Gene expression profiling defined two expression
194 JAK-STAT Signaling in Diseases

signatures in DLBCL: Germ cell B (GCB-DLBCL) or activated B-cell (ABC-DLBCL) type, the
latter conferring a worse prognosis. In children, the GCB-DLBCL subtype is more common and
is not associated with prognosis. A third NHL subtype is the primary mediastinal large B-cell
lymphoma (PMBCL), mainly occurring in young adult women with a distinct gene expression
pattern, treatment and outcome. Anaplastic large-cell lymphoma (ALCL) is the main peripheral
T-cell lymphoma in the pediatric age group and comprises 10–15% of NHL. In over 95% of
pediatric ALCL, the ALK, Anaplastic Lymphoma Kinase, gene rearrangement is identified and
drives lymphomagenesis. The ALK-negative (ALK-) ALCL has a different gene expression profile
and a worse prognosis.
Due to the higher incidence of NHL in adults, most of molecular studies presented here were
performed in adults only or using mixed pediatric/adult samples. We will also discuss the
relevance of these findings in the pediatric population when applicable.

11.4.1 JAK-STAT Pathway Gene Alterations Are Frequent in Different


Lymphoma Subtypes
11.4.1.1 JAK1 Alterations
JAK1 mutations occur mostly in T-cell lymphomas (Table 11.3). They represent 15% of 88 ALCL,
6% of 110 natural killer-cell/T-cell lymphoma (NKTCL) and 11% of 9 cutaneous T-cell
lymphoma (CTCL) cases (samples include both children and adults) (Crescenzo et al. 2015;
Song et al. 2018). In ALCL, JAK1 mutations are located in the JH1 kinase domain and involve
almost exclusively the G1097 residue with several missense mutations substituting to aspartic acid,
asparagine or serine (Crescenzo et al. 2015). In vivo, mutant JAK1 leads to increased levels of
pSTAT3 and cell proliferation. This oncogenic mutation is synergistic with STAT3 mutations and
exhibits sensitivity to JAK inhibition (Crescenzo et al. 2015).

11.4.1.2 JAK2 Alterations


JAK2 sequence mutations have not been described in lymphomas. In contrast to MPN, no
canonical JAK2V617F mutations were found in cell lines derived from lymphomas, PMBCL and
cHL (Melzner et al. 2006; Scott et al. 2005; Wu et al. 2009). Moreover, activating JAK2 mutations
in exon 12, described in MPN, were not found in PBMCL (Wu et al. 2009). JAK2 fusions have
been identified in 4 adult patients with cHL (Roosbroeck et al. 2011). Two patients carry
a SEC31A-JAK2 fusion resulting from a reciprocal t(4;9)(q21;p24) translocation. Its expression
in hematopoietic cell lines results in cytokine-independent growth and can be abolished by
pharmacological JAK inhibition. SEC31A-JAK2–transduced murine cells lead to a fatal T-ALL
in 6 out of 16 mice. Two additional patients harbor JAK2 fusions but the respective partners
could not be identified.

11.4.1.3 JAK3 Alterations


The identification of somatic JAK3 mutations has been made in NKTCL, representing
20-35% of the studied cohort (adult/pediatric) and in approximately 20% of CTCL samples
(Almeida et al. 2015; Bouchekioua et al. 2014; Koo et al. 2012; McGirt et al. 2015).
NKTCL has the highest prevalence in Asia and Latin America but remains extremely rare in
children among whom only a few case reports have been reported (Lai and Dunleavy 2013; Pillai
et al. 2016).
Two recurrent missense cytosine-for-thymine substitutions (A572V and A573V) were described
in NHL. Both are located in exon 12 of JAK3, within the JH2 domain. These mutations are
mostly detected at the heterozygous state but some lymphomas harbor compound heterozygous or
homozygous mutations. A single patient was reported with a V722I mutation, located in exon 16,
also belonging to the JH2 domain (Bouchekioua et al. 2014). The JAK3A572V mutation has also
JAK-STAT Signaling in Hematologic Malignancies 195

TABLE 11.3
JAK-STAT Alterations and Targeted Therapies in Lymphomas
Treatment options
JAK-STAT
Lymphomas alterations Frequency Pre-clinical Clinical

Hodgkin Lymphoma
JAK2 fusion Individual cases Type I JAK inhibitors Ruxolitinib in adult
(n=4) patients, phase II
STAT6 mutations Around 30% of cases completed
9p24 amplification Up to 90% of cases (NCT01877005) and
ongoing
SOCS1 mutation Around 55% of cases
(NCT01965119).
PTPN1 mutation Around 20% of cases Pacritinib in adults,
phase I completed
(NCT00741871)

B-cell non-Hodgkin lymphoma


DLBCL STAT6 mutations Around 15% of adults Type I JAK inhibitors Pacritinib in adults,
GCB-DLBCL phase I completed
SOCS1 mutation Around 27% of adults (NCT00741871)
MYD88 mutations Around 30% of
ABC-DLBCL
PMBCL 9p24 amplification Around 35% of cases Type I JAK inhibitors
SOCS1 mutation Around 55% of cases
PTPN1 mutation Around 20% of cases
STAT6 mutations 36-72% of cases

T-cell non-Hodgkin lymphoma


ALCL JAK1 mutation Around 15% of cases Type I JAK inhibitors Pacritinib in adults,
TYK2 activation Most of ALCL TYK2 inhibitors phase I completed
(NCT00741871).
STAT3 mutation Around 10% of Type I JAK inhibitors
ALK- ALCL Cerdulatinib in adult
patients, phase IIa
NKTCL JAK1 mutation Around 6% of cases Type I JAK inhibitors (NCT01994382)
JAK3 mutation 20-35% of cases JAK3 inhibitors
STAT3 mutation Around 20% of cases Type I JAK inhibitors
PTCL PTPN2 mutation Individual cases Type I JAK inhibitors
(n=2)
CTCL JAK1 mutation Around 10% of cases Type I JAK inhibitors
JAK3 mutation Around 20% of cases JAK3 inhibitors

DLBCL: Diffuse large B-cell lymphoma, PMBCL: Primary mediastinal large B-cell lymphoma, ALCL: Anaplastic
large-cell lymphoma, NKTCL: natural killer-cell/T-cell lymphoma, PTCL: Peripheral T-cell lymphoma, CTCL:
cutaneous T-cell lymphoma

been reported in adult patients with CTCL (Cornejo et al. 2009; McGirt et al. 2015). In NKTCL
cell lines, JAK3A572V leads to constitutive activation of both JAK3 and STAT5, resulting in
cytokine-independent growth (Koo et al. 2012). JAK3A572V–transduced cells in a murine model
lead to a fatal TCRαβ+ T-lymphoproliferative disease in all mice (Cornejo et al. 2009). Mutations
in the FERM domain of JAK3 have been reported in adult T-cell leukemia/lymphoma, a very rare
disease not occurring during childhood.
196 JAK-STAT Signaling in Diseases

11.4.1.4 TYK2 Alterations


To date, TYK2 mutations have never been reported in lymphomas. However, activation of TYK2
has been identified as a central event in ALK+ and ALK- ALCL (Prutsch et al. 2018). TYK2 is
highly expressed and its knockdown or inhibition in human ALCL cells induces death. In
transgenic mice models of ALCL, the suppression of TYK2 reduces tumoral growth. The effect
on cell growth depends on STAT1 and STAT3 phosphorylation and is strengthened by an
autocrine mechanism involving IL-10 and IL-22: TYK2 overexpression results in production of
these cytokines which bind to their cognate receptor on tumor cells, activating STAT1 and STAT3
(Prutsch et al. 2018). In ALK- ALCL with a JAK-STAT mutation, the presence of cytokine
signaling is also necessary for cell survival (Chen et al. 2017). These data suggest that cytokine
signaling plays a major role in sustaining cell viability in spite of constitutive activation of the
JAK-STAT pathway. The molecular mechanism underlying TYK2 activation in ALK+ ALCL is
not known; however, in some patients with ALK- ALCL, transcriptomic analysis unravels
recurrent NPM1-TYK2 or NF-κB2-TYK2 fusions (Crescenzo et al. 2015; Velusamy et al. 2014).
NPM1-TYK2 fusion is found in cutaneous ALK- ALCL and is also present in lymphomatoid
papulosis. A single case with ALK- ALCL harboring a PABPC4-TYK2 fusion was reported
(Crescenzo et al. 2015). The NPM1-TYK2 fusion gene induces constitutive activation of TYK2
and leads to phosphorylation of STAT1, STAT3 and STAT5. TYK2 knockdown in NPM1-TYK2
MyLa cell lines results in a significant decrease in cell proliferation (Velusamy et al. 2014). As for
the NF-κB2-TYK2 fusion, it constitutes a chimeric transcription factor (Crescenzo et al. 2015). Its
transduction in cell lines results in tumorigenesis and induces JAK and STAT3 phosphorylation.
Another fusion of NF-κB2 within the intracytoplasmic domain of ROS1 (NF-κB2-ROS1) has
been identified in the same work (Crescenzo et al. 2015). ROS1 is a JAK-STAT-independent
tyrosine kinase with structural similarities to the ALK protein. Interestingly, the effect of the
fusion phenocopies that of NF-κB2-TYK2 in regards to STAT3 dependency.

11.4.1.5 STAT3 Alterations


Somatic STAT3 mutations were primarily described in granular lymphocytic leukemia but have
been found in 21% of 101 patients with NKTCL, 10% of 88 patients with ALK- ALCL and 3% of
44 patients with cutaneous ALCL (mixed adult/pediatric samples) (Crescenzo et al. 2015; Hu
et al. 2013; Koskela et al. 2012; Lohr et al. 2012; Morin et al. 2011; Song et al. 2018). These point
mutations are not recurrent but occur predominantly within the SH2 domain. The pathogenicity
of each mutation is not equivalent. In murine lymphoid Ba/F3 cells, among 4 mutations predicted
to be pathogenic (p.D427H, p.E616G, p.E616K and p.E696K), all four lead to STAT3 phosphor­
ylation, but only the p.E616K mutation can confer cytokine-independent growth (Song et al.
2018). In vitro, STAT3Y640F cell lines are constitutively activated and can be abrogated by STAT3
inhibition. STAT3Y640F-transplanted hematopoietic cells in mice lead to the development of
a myeloproliferative-like disease (Couronné et al. 2013).
In NKTCL cell lines, pSTAT3 levels correlate with PD-L1 expression on the tumor cell surface
(Song et al. 2018). This effect is mediated by STAT3 direct regulation of PD-L1 transcriptional
activity. Remarkably, STAT3 silencing and inhibition decrease PD-L1 expression, providing
a potential therapeutic approach combining STAT3 inhibitors to anti-PD-1/PD-L1 antibodies
(Song et al. 2018).
In T-cell lymphoma, STAT3 mutations co-exist with other JAK-STAT mutations in 62% of 171
cases (Song et al. 2018). As previously mentioned, the co-existence of STAT3 and JAK1 mutations in
ALK- ALCL exerts synergistic effect on cell proliferation in vitro (Crescenzo et al. 2015).

11.4.1.6 STAT6 Alterations


Heterozygous STAT6 mutations have been found in 36-72% of PMBCL, 14% of adult GCB­
DLBCL and 32% of adult/pediatric cHL (Dubois et al. 2016; Ritz et al. 2009; Tiacci et al. 2018).
JAK-STAT Signaling in Hematologic Malignancies 197

These mutations occur in the hotspot DNA-binding domain within the exon 12 and overlap between
the different disease subtypes. Interestingly, a large proportion of patients harbor concomitant
alterations in other genes regulating the JAK-STAT pathway, such as SOCS1 and JAK2, providing
additional evidence to support that alterations affecting this pathway are not mutually exclusive.
Nevertheless, the downstream effect of STAT6 mutations is not fully understood. They could lead to
constitutive phosphorylation resulting in tumor growth (Ritz et al. 2009; Tiacci et al. 2018); however,
mutant-transduced cell lines exhibit an impaired rather than enhanced DNA binding activity of
STAT6 (Ritz et al. 2009). Moreover, in a limited gene expression study, the transcription of STAT6
target genes was not modulated by these mutations (Ritz et al. 2009).
STAT6 mutations have also been described in adult follicular lymphoma (Yildiz et al. 2015).
However, this entity is quite uncommon in children and exhibits distinct pathological and genetic
features from the adult counterpart (Lorsbach et al. 2002; Oschlies et al. 2006). Notably, MAPK
pathway mutations are more prevalent and the involvement of STAT6 mutations in these
pediatric cases has not been reported (Louissaint et al. 2016). The clinical relevance of these
findings is currently not known in children.

11.4.2 JAK-STAT Pathway Collaborates in Lymphomagenesis through the Recurrent 9p24


Amplification in cHL and PMBCL
PMBCL gene expression studies have revealed that it displays a distinct profile compared to that
of DLBCL and cHL (Rosenwald et al. 2003; Savage et al. 2003). Recurrent gene alterations are
shared between these different B-cell lymphomas, especially the classical focal 9p24.1 amplifica­
tion (Borchmann and Engert 2017; Lenz et al. 2008; Meier et al. 2009). Up to 94% of patients
with cHL may harbor a 9p24.1 copy gain (> 2 copies) or amplification (>5 copies) (Roemer et al.
2016). The intensity and the spanning of the amplified region vary, but the minimal common
region of copy number gain seems to encompass 7–10 genes among which figure PD-L1, PD-L2,
KDM4C, RANBP6 and JAK2. Being a part of the amplicon, JAK2 may contribute to the
pathogenesis and mediate several effects in lymphomas. In cHL and PMBCL cell lines, JAK2
knockdown or drug-mediated inhibition induces apoptosis but does not inhibit proliferation (Rui
et al. 2010). The effect of JAK2 on cell viability is dependent on IL-13, of which its secretion is
upregulated by an autocrine loop (Rui et al. 2010). Moreover, JAK2 has a direct effect on
increasing PD-L1 expression (Green et al. 2010). Since JAK2 is not the only upregulated gene
within the amplicon, the other genes located in the amplicon probably share an important role in
tumorigenesis. In fact, in contrast to JAK2, KDM4C and RANBP6 knockdown reduces prolifera­
tion without inducing apoptosis (Rui et al. 2010). Moreover, the cooperation within the amplified
region fosters cell survival and proliferation. Functional analyses showed that KDM4C, a histone
demethylase, cooperates with JAK2 in the epigenome modification (Rui et al. 2010). The
inhibition of both KDM4C and JAK2 increases heterochromatin and kills PMBCL and HL cell
lines (Rui et al. 2010). Upon action of these genes in cHL, transcription–repression of hundred
genes is reduced, several of which are likely to be involved in lymphomagenesis. One of them is
MYC, which does not allow cell proliferation by itself but remains important in promoting
survival (Rui et al. 2010). In conclusion, this amplicon provides multiple cooperative key genes
involving the JAK-STAT pathway and providing a potential therapeutic action, notably through
JAK inhibition in combination with immune checkpoint inhibitors, as demonstrated in NKTCL
(Song et al. 2018).

11.4.3 Loss-of-Function of JAK-STAT Regulators Is Frequent in Lymphomas


11.4.3.1 SOCS1 Mutations
SOCS1 mutations have been reported as a mechanism of JAK-STAT pathway activation in some
lymphomas. Suppressor-of-cytokine-signaling (SOCS) family proteins constitute major regulators of
JAK activity and are involved in cellular homeostasis. These proteins contain a SH2 domain, which
198 JAK-STAT Signaling in Diseases

binds to JAK, and ultimately lead to proteosomal degradation through poly-ubiquitination


(Ungureanu et al. 2002). STAT5 activation induces SOCS1 transcription, which limits JAK
activity (Hennighausen and Robinson 2008). Therefore, loss-of-function mutations in SOCS1
result in the abolition of this negative-feedback loop and hyperactivation of the JAK-STAT pathway.
SOCS1 mutations have been identified in several types of B-cell lymphomas in pediatric and adult
patients (Melzner et al. 2005; Mottok et al. 2009; Weniger et al. 2006). They are present in up to 56%
of cHL (pediatric/adult samples), 55% of PMBCL and 27% of adult DLBCL (Dubois et al. 2016;
Mottok et al. 2009; Tiacci et al. 2018; Weniger et al. 2006). Although less frequent, 4% of 171 T-cell
lymphomas harbor a SOCS1 mutation (Song et al. 2018). In B-cell lymphomas, these mutations
result from the somatic hypermutation process occurring during mature B-cell development upon
B-cell receptor engagement (Mottok et al. 2009, 2007). Consequently, there is no specific hotspot in
SOCS1 and SOCS1 alterations translate in most cases in partial or complete deletion of the protein.
Most frequent types of alterations are non-sense mutations resulting in a truncated protein followed
by interstitial deletions (Mottok et al. 2007). Both heterozygous and homozygous mutations have
been reported (Melzner et al. 2005; Mottok et al. 2009). In PMBCL-derived MedB-1 cell lines,
impaired SOCS1 leads to high levels of pSTAT5 and pJAK2 (Melzner et al. 2005). The constitutive
JAK2 activation was due to its defective degradation, which can be corrected by ectopic SOCS1
expression. In this cell line, SOCS1 loss-of-function also mediates the loss of RB1 expression,
another important tumor-suppressor gene (Melzner et al. 2005). SOCS1 mutations and 9p24.1
amplification are not mutually exclusive, since a subset of patients with PBMCL harbors both
alterations (Melzner et al. 2005; Mottok et al. 2007). Furthermore, some cHL patients have been
reported with both SOCS1 and STAT6 mutations (Tiacci et al. 2018).

11.4.3.2 PTPN1 Mutations


Another mechanism of regulation of JAK-STAT activation is the dephosphorylation of activated
JAK/TYK2 proteins. PTPN1 and PTPN2 are phosphatases removing the phosphorylation of
activated JAK2/TYK2 and JAK1/JAK3, respectively. Whole-genome and whole-transcriptome
sequencing identified somatic PTPN1 mutations in 20% of 6 cHL and 22% of 77 PMBCL cases
(Gunawardana et al. 2014). Mutation types include predominantly missense, then nonsense and
frameshift mutations. Most mutations are located in the catalytic domain and the impairment impact
varies according to the mutation. In cHL cell line, these mutations decrease PTPN1 phosphatase
activity, leading to hyperphosphorylation of JAK-STAT proteins. Co-occurrence of PTPN1 and
SOCS1 mutations were frequent in this cohort (Gunawardana et al. 2014). Functional analysis of
PTP1BΔ6, a PTPN1 variant lacking exon 6, demonstrates that its expression in cHL cell lines
induces STAT activity and cell proliferation. This dominant negative variant protects against
cytotoxic agents and impedes the normal PTP1B function by heterodimerization (Zahn et al. 2017).

11.4.3.3 PTPN2 Mutations


Bi-allelic inactivation of PTPN2 has been found in 2 out of 39 patients with peripheral T-cell lymphoma
not otherwise specified (Kleppe et al. 2011). One patient carries two frameshift mutations (p.G41fs and
p.A45fs) and the second patient has a missense mutation (c.133G>A) in one allele and a 18p deletion in
the other allele. One variant was found in a patient with cHL but without functional consequence. It has
been shown that cHL SUP-HD1 cell lines have very low PTPN2 protein levels. These cells carry
a missense mutation (p.L249P) in one allele and a gene deletion in the other. In a similar fashion to
PTPN1 loss-of-function, JAK-STAT proteins are constitutively activated when PTPN2 is inactivated.

11.4.4 MYD88 Mutations Activate the JAK-STAT Pathway in ABC-DLBCL


ABC-DLBCL is characterized by a constitutive activation of the anti-apoptotic NF-κB signaling
pathway (Compagno et al. 2009; Davis et al. 2001). Mutations in MYD88 have been reported in
JAK-STAT Signaling in Hematologic Malignancies 199

approximately 30% of ABC-DLBCL and 10% of mucosa-associated lymphoid tissue lymphoma


(Ngo et al. 2011; Schmitz et al. 2018). They are also present in more than 90% of Waldenstrom’s
macroglobulinemia, a disease restricted only to adult patients (Treon et al. 2012).
MYD88 is a signaling adaptor protein downstream of IL-1, IL-18 and all toll-like receptors that
activate the NF-κB pathway upon ligand binding. The large majority of mutations identified
consists of L265P point mutations, located in the Toll-like/IL-1 receptor domain (Ngo et al. 2011).
In ABC-DLBCL cell line, MYD88L265P results in a gain-of-function and activates NF-κB signaling
pathway. In addition, these cells exhibit activated STAT3, which synergizes with NF-κB to promote
cell proliferation and survival (Ding et al. 2008; Lam et al. 2008). These two pathways interact
through an autocrine loop involving IL-6 and IL-10 (Gupta et al. 2012). Through IRAK1 and
IRAK4 signaling, NF-κB stimulates IL-6, IL-10 and interferon-β synthesis. These cytokines activate
their receptors on the surface of lymphoma cell, causing STAT3 activation, which will enhance
NF-κB activation (Ngo et al. 2011).

11.4.5 Therapeutic Implications of the JAK-STAT Pathway in Lymphomas


Due to the frequent activation of the JAK-STAT pathway in lymphomas and its central role in
lymphomagenesis, the use of JAK-STAT pathway inhibitors seems like a promising therapeutic
approach but its translation into the clinic was not as successful as expected. Only three drugs
have moved to early phase trials for adult lymphoma patients.
First, ruxolitinib has demonstrated in vitro and in vivo anti-proliferative activity (Lee et al.
2018). A phase II study (NCT01877005) of ruxolitinib in patients with relapsed or refractory HL
reported modest responses, while another multicentric study (NCT01965119) is ongoing (Kim
et al. 2014; Neste et al. 2018). Pacritinib (SB1518) demonstrated some response in patients with
cHL and various NHL in a phase I study (NCT00741871) (Hart et al. 2011; Younes et al. 2012).
Cerdulatinib has shown in vitro activity in DLBCL cell lines and allowed responses in around
40% of T-cell lymphoma in the preliminary analysis of the phase IIa trial (NCT01994382)
(Hamlin et al. 2018; Ishikawa et al. 2018; Ma et al. 2015).
Other JAK–family inhibitors have been tested clinically but their development has since halted
as in the examples of fedratinib because of Wernicke encephalopathy and lestaurtinib for lack of
efficacy (Diaz et al. 2011; Hao et al. 2014; Marshall et al. 2005). Some agents have only been used
in preclinical studies (tofacitinib, PRN371 and AZD1480) (Bouchekioua et al. 2014; Derenzini
et al. 2011; Koo et al. 2012; Nairismägi et al. 2018).
As discussed above, there is biologic rationale for the combination of a JAK-STAT inhibitor with
other novel agents, notably with immune-checkpoint inhibitors, but no such experience has been
reported in lymphoma yet (only a case report related the efficacy of the combination between nivolumab
and ruxolitinib in a patient with a Merkel cell carcinoma) (Debureaux et al. 2018). Nevertheless, the
combination of ruxolitinib with the MMAE-conjugated anti-CD30 monoclonal antibody brentuximab­
vedotin has demonstrated an increased efficacy in vitro and in vivo (Ju et al. 2016).

11.4.6 Conclusion
JAK-STAT hyperactivation is common in lymphomas and can be dysregulated in numerous subtypes
by a diverse array of genomic alterations. This pathway plays a central role in lymphomagenesis by
promoting cell survival and proliferation. Furthermore, several examples in different lymphoma
subtypes illustrate that the JAK-STAT pathway was not the only deregulated signal but interacts
with other signaling pathways to promote cell proliferation and/or immune-escape.
Due to frequency of the JAK-STAT hyperactivation and its important effect in lymphomas, its
inhibition seems like a logical strategy. Despite a strong rationale derived from robust in vitro and
ex vivo data, clinical development of JAK inhibitors in this patient population is lagging behind
because of the lack of durable response in early phase trials. Perhaps a more effective approach
would be the combination of JAK-STAT pathway inhibition with cytotoxic chemotherapy or
200 JAK-STAT Signaling in Diseases

immunotherapeutic agents such as brentuximab vedotin and anti-PD1/PD-L1 antibodies, given


that JAK-STAT pathway is not the only driver present in lymphoma cells residing within a rich
immune microenvironment.

11.5 Summary
JAK-STAT dysregulation is the recurrent central event upon the different hematologic malignan­
cies reviewed herein. The underlying mechanisms activating JAK-STAT signaling are diverse and
differ between disease types. JAK-STAT family of genes may be directed altered but other genes
regulating this pathway might be also affected. Not all of the genomic alterations identified are
functionally characterized, but those which have been studied all converge to the central theme of
reinforcing JAK-STAT activity. Altogether, these data attribute this pathway as the cornerstone of
tumorigenesis. The collaborating events foster JAK-STAT signaling which drives cell proliferation
in the different models presented. In most of in vitro and in vivo experiments, this proliferation
could have been attenuated or even abolished with JAK inhibitors.
Based on these compelling data, the clinical use of type I JAK inhibitor ruxolitinib has been
swiftly increased in the last 10 years; first in MPN then expanding to almost all hematologic
malignancies. In most cases, ruxolitinib results in symptomatic improvement but survival benefits
have not been convincingly demonstrated. In fact, some studies suggest that ruxolitinib is unable
to eliminate the primitive clone and acquired mechanisms of resistance have emerged from
longterm drug exposure. These phenomena may be in part explained by the fact that the JAK­
STAT pathway is undoubtedly involved in a large number of hematologic malignancies to
promote cell proliferation, but it remains unclear if it constitutes the driver event. Hence, even in
MPN where JAK-STAT involvement has been the most studied, the role of JAK2V617F in
initiating the disease remains controversial.
The selection of the ideal JAK inhibitor to use in the clinic is yet unknown. Currently, only type
I JAK inhibitors are being clinically investigated. They have an acceptable safety profile with
moderate immunosuppression as the main side effect. The use of type II JAK inhibitors leads to
stronger JAK inhibition but potentially more adverse toxicities as well. In particular, since JAK­
STAT signaling plays a key role in the immune system, its inhibition represents an effective
therapeutic approach in several inflammatory and auto-immune diseases (Banerjee et al. 2017).
This pathway inhibition affects T-cells, NK-cells and dendritic cells’ function (Assi et al. 2014;
Curran et al. 2017; Keohane et al. 2015; Parampalli Yajnanarayana et al. 2015). It might be of
benefit to limit the cancer-associated inflammation, as it seems to be in MPN. (Hasselbalch 2012;
Tefferi et al. 2011) However, it also dampens the tumor immunosurveillance and the potential
action of the immune system upon cell proliferation (Groner and von Manstein 2017). Therefore,
the overall effect of JAK inhibition depends on the balance between its tumoral and immunolo­
gical effects. The benefits are probably variable depending on the specific tumor type and the
given host. The current data on JAK-STAT implication in the micro-environment and immune
reaction surrounding tumor cells is yet largely unknown. Future research should focus on an
integrative approach to study genomic alterations and their functional impact on the tumor cells
and their microenvironment. This will allow a deepened understanding of the disease biology to
design more effective therapeutic strategies.

REFERENCES
Al-Ali, Haifa, Martin Griesshammer Kathrin, Philipp le Coutre, Cornelius F. Waller, Anna
Marina Liberati, Philippe Schafhausen, Renato Tavares, et al. 2016. “Safety and Efficacy of
Ruxolitinib in an Open-Label, Multicenter, Single-Arm Phase 3b Expanded-Access Study in
Patients with Myelofibrosis: A Snapshot of 1144 Patients in the JUMP Trial.” Haematologica 101
(9): 1065–73. doi: 10.3324/haematol.2016.143677.
JAK-STAT Signaling in Hematologic Malignancies 201

Almeida, Silva, Ana Carolina Da, Francesco Abate, Hossein Khiabanian, Estela Martinez-Escala,
Joan Guitart, Cornelis P. Tensen, Maarten H. Vermeer, Raul Rabadan, Adolfo Ferrando, and
Teresa Palomero. 2015. “The Mutational Landscape of Cutaneous T Cell Lymphoma and Sézary
Syndrome.” Nature Genetics 47 (12): 1465–70. doi: 10.1038/ng.3442.
An, Wenbin, Yang Wan, Ye Guo, Xiaojuan Chen, Yuanyuan Ren, Jingliao Zhang, Lixian Chang,
Wei Wei, Peihong Zhang, and Xiaofan Zhu. 2014. “CALR Mutation Screening in Pediatric
Primary Myelofibrosis.” Pediatric Blood & Cancer 61 (12): 2256–62. doi: 10.1002/pbc.25211.
Anand, Shubha, Frances Stedham, Philip Beer, Emma Gudgin, Christina A. Ortmann, Anthony Bench,
Wendy Erber, Anthony R. Green, and Brian J. P. Huntly. 2011. “Effects of the JAK2 Mutation on
the Hematopoietic Stem and Progenitor Compartment in Human Myeloproliferative Neoplasms.”
Blood 118 (1): 177–81. doi: 10.1182/blood-2010-12-327593.
Ang, Sonny O., Hua Chen, Victor R. Kiichi Hirota, Jaroslav Jelinek Gordeuk, Yongli Guan, Enli Liu,
et al. 2002. “Disruption of Oxygen Homeostasis Underlies Congenital Chuvash Polycythemia.”
Nature Genetics 32 (4): 614–21. doi: 10.1038/ng1019.
Araki, Marito, Yinjie Yang, Nami Masubuchi, Yumi Hironaka, Hiraku Takei, Soji Morishita,
Yoshihisa Mizukami, et al. 2016. “Activation of the Thrombopoietin Receptor by Mutant Calreti­
culin in CALR-Mutant Myeloproliferative Neoplasms.” Blood 127 (10): 1307–16. doi: 10.1182/
blood-2015-09-671172.
Arber, Daniel A., Attilio Orazi, Robert Hasserjian, Michael J. Jürgen Thiele, Michelle M. Borowitz, Clara
D. Le Beau, Mario Cazzola Bloomfield, and James W. Vardiman. 2016. “The 2016 Revision to the
World Health Organization Classification of Myeloid Neoplasms and Acute Leukemia.” Blood 127
(20): 2391–405. doi: 10.1182/blood-2016-03-643544.
Aricò, Maurizio, Stephen P. Martin Schrappe, William L. Hunger, Valentino Conter Carroll,
Stefania Galimberti, Atsushi Manabe, et al. 2010. “Clinical Outcome of Children with Newly
Diagnosed Philadelphia Chromosome–Positive Acute Lymphoblastic Leukemia Treated
between 1995 and 2005.” Journal of Clinical Oncology 28 (31): 4755–61. doi: 10.1200/
JCO.2010.30.1325.
Armstrong, Scott A., Meghann E. Mabon, Lewis B. Silverman, Aihong Li, John G. Gribben, Edward
A. Fox, Stephen E. Sallan, and Stanley J. Korsmeyer. 2004. “FLT3 Mutations in Childhood Acute
Lymphoblastic Leukemia.” Blood 103 (9): 3544–6. doi: 10.1182/blood-2003-07-2441.
Assi, Hikmat, Jaclyn Espinosa, Sarah Suprise, Michael Sofroniew, Robert Doherty, Daniel Zamler, Pedro
R. Lowenstein, and Maria G. Castro. 2014. “Assessing the Role of STAT3 in DC Differentiation
and Autologous DC Immunotherapy in Mouse Models of GBM.” PLoS One 9 (5): e96318. doi:
10.1371/journal.pone.0096318.
Banerjee, Shubhasree, Ann Biehl, Massimo Gadina, Sarfaraz Hasni, and Daniella M. Schwartz. 2017.
“JAK–STAT Signaling as a Target for Inflammatory and Autoimmune Diseases: Current and
Future Prospects.” Drugs 77 (5): 521–46. doi: 10.1007/s40265-017-0701-9.
Barbui, T., J. Thiele, A. M. Vannucchi, and A. Tefferi. 2014. “Rethinking the Diagnostic Criteria of
Polycythemia Vera.” Leukemia 28 (6): 1191–5. doi: 10.1038/leu.2013.380.
Bartalucci, Niccolò, Lorenzo Tozzi, Costanza Bogani, Serena Martinelli, Giada Rotunno, Jean-Luc
Villeval, and Alessandro M. Vannucchi. 2013. “Co-Targeting the PI3K/MTOR and JAK2 Signal-
ling Pathways Produces Synergistic Activity against Myeloproliferative Neoplasms.” Journal of
Cellular and Molecular Medicine 17 (11): 1385–96. doi: 10.1111/jcmm.12162.
Baughn, L. B., M. M. Meredith, L. Oseth, T. A. Smolarek, and B. Hirsch. 2018. “SH2B3 Aberrations
Enriched in IAMP21 B Lymphoblastic Leukemia.” Cancer Genetics 226–227 (October): 30–5. doi:
10.1016/j.cancergen.2018.05.004.
Baxter, E. Joanna, Linda M Scott, Peter J Campbell, Clare East, Nasios Fourouclas, Soheila Swanton,
George S Vassiliou, et al. 2005. “Acquired Mutation of the Tyrosine Kinase JAK2 in Human
Myeloproliferative Disorders.” The Lancet 365 (9464): 1054–61. doi: 10.1016/S0140-6736(05)71142-9.
Beer, Philip A., Peter J. Campbell, Linda M. Scott, Anthony J. Bench, Wendy N. Erber, David Bareford,
Bridget S. Wilkins, et al. 2008. “MPL Mutations in Myeloproliferative Disorders: Analysis of the
PT-1 Cohort.” Blood 112 (1): 141–9. doi: 10.1182/blood-2008-01-131664.
Ding, B. Belinda, J. Jessica Yu, Raymond Y.-L. Yu, Lourdes M. Mendez, Rita Shaknovich, Yonghui Zhang,
Giorgio Cattoretti, and B. Hilda Ye. 2008. “Constitutively Activated STAT3 Promotes Cell
202 JAK-STAT Signaling in Diseases

Proliferation and Survival in the Activated B-Cell Subtype of Diffuse Large B-Cell Lymphomas.”
Blood 111 (3): 1515–23. doi: 10.1182/blood-2007-04-087734.
Bellanné-Chantelot, Christine, Isabelle Chaumarel, Myriam Labopin, Florence Bellanger,
Véronique Barbu, Claudia De Toma, François Delhommeau, et al. 2006. “Genetic and Clinical
Implications of the Val617Phe JAK2 Mutation in 72 Families with Myeloproliferative Disorders.”
Blood 108 (1): 346–52. doi: 10.1182/blood-2005-12-4852.
Benekli, Mustafa, Kathleen A. Zheng Xia, Laurie A. Donohue, Lynda A. Ford, Maria R. Pixley, Heinz
Baumann Baer, and Meir Wetzler. 2002. “Constitutive Activity of Signal Transducer and Activator
of Transcription 3 Protein in Acute Myeloid Leukemia Blasts Is Associated with Short Disease-Free
Survival.” Blood 99 (1): 252–7. doi: 10.1182/blood.V99.1.252.
Bercovich, Dani, Ithamar Ganmore, Linda M Scott, Gilad Wainreb, Yehudit Birger,
Arava Elimelech, Chen Shochat, et al. 2008. “Mutations of JAK2 in Acute Lymphoblastic
Leukaemias Associated with Down’s Syndrome.” The Lancet 372 (9648): 1484–92. doi:
10.1016/S0140-6736(08)61341-0.
Biondi, Andrea, Martin Schrappe, Paola De Lorenzo, Anders Castor, Giovanna Lucchini,
Virginie Gandemer, Rob Pieters, et al. 2012. “Imatinib after Induction for Treatment of Children
and Adolescents with Philadelphia-Chromosome-Positive Acute Lymphoblastic Leukaemia
(Esphall): A Randomised, Open-Label, Intergroup Study.” The Lancet Oncology 13 (9): 936–45.
doi: 10.1016/S1470-2045(12)70377-7.
Bock, Charles E. de, Sandrine Degryse, Sofie Demeyer, Bram Sweron, Olga Gielen, Nicole Mentens,
Ellen Geerdens, and Jan Cools. 2014. “Synergism between HOXA9 and Mutant JAK3 (M511I)
Leads to Rapid Leukemia Development within an In Vivo Murine Bone Marrow Transplant
Model.” Blood 124 (21): 1078. www.bloodjournal.org/content/124/21/1078.
Bodegom, Diederik van, Jun Zhong, Nadja Kopp, Chaitali Dutta, Min-Sik Kim, Liat Bird,
Oliver Weigert, et al. 2012. “Differences in Signaling through the B-Cell Leukemia Oncoprotein
CRLF2 in Response to TSLP and through Mutant JAK2.” Blood 120 (14): 2853–63. doi: 10.1182/
blood-2012-02-413252.
Bolouri, Hamid, Jason E. Farrar, Rhonda E. Timothy Triche, Emilia L. Ries, Todd A. Lim, Yussanne
Ma Alonzo, et al. 2018. “The Molecular Landscape of Pediatric Acute Myeloid Leukemia Reveals
Recurrent Structural Alterations and Age-Specific Mutational Interactions.” Nature Medicine 24
(1): 103–12. doi: 10.1038/nm.4439.
Borchmann, Sven, and Andreas Engert. 2017. “The Genetics of Hodgkin Lymphoma: An Overview
and Clinical Implications.” Current Opinion in Oncology 29 (5): 307–14. doi: 10.1097/CCO.
0000000000000396.
Bose, Prithviraj, and Srdan Verstovsek. 2017. “JAK2 Inhibitors for Myeloproliferative Neoplasms: What
Is Next?” Blood 130 (2): 115–25. doi: 10.1182/blood-2017-04-742288.
Bouchekioua, A., L. Scourzic, O. de Wever, Y. Zhang, P. Cervera, A. Aline-Fardin, T. Mercher, et al.
2014. “JAK3 Deregulation by Activating Mutations Confers Invasive Growth Advantage in
Extranodal Nasal-Type Natural Killer Cell Lymphoma.” Leukemia 28 (2): 338–48. doi: 10.1038/
leu.2013.157.
Bousfiha, Aziz, Leïla Jeddane, Capucine Picard, H. Fatima Ailal, Bobby Gaspar, Waleed Al-Herz,
Talal Chatila, et al. 2018. “The 2017 IUIS Phenotypic Classification for Primary Immunodeficiencies.”
Journal of Clinical Immunology 38 (1): 129–43. doi: 10.1007/s10875-017-0465-8.
Buitenkamp, Trudy D., Shai Izraeli, Martin Zimmermann, Erik Forestier, Nyla A. Heerema,
Marry M. van Den Heuvel-Eibrink, Rob Pieters, et al. 2014. “Acute Lymphoblastic Leukemia in
Children with down Syndrome: A Retrospective Analysis from the Ponte Di Legno Study Group.”
Blood 123 (1): 70–7. doi: 10.1182/blood-2013-06-509463.
Burkhardt, Birgit, Martin Zimmermann, Ilske Oschlies, Felix Niggli, Georg Mann, Reza Parwaresch,
Hansjoerg Riehm, Martin Schrappe, and Alfred Reiter. 2005. “The Impact of Age and Gender on
Biology, Clinical Features and Treatment Outcome of Non-Hodgkin Lymphoma in Childhood and
Adolescence.” British Journal of Haematology 131 (1): 39–49. doi: 10.1111/j.1365-2141.2005.05735.x.
Butcher, C. M., U. Hahn, L. B. To, J. Gecz, E. J. Wilkins, H. S. Scott, P. G. Bardy, and R. J. D’Andrea.
2008. “Two Novel JAK2 Exon 12 Mutations in JAK2V617F-Negative Polycythaemia Vera
Patients.” Leukemia 22 (4): 870–3. doi: 10.1038/sj.leu.2404971.
JAK-STAT Signaling in Hematologic Malignancies 203

Cabagnols, Xénia, Fabrizia Favale, Florence Pasquier, Kahia Messaoudi, Jean Philippe Defour, Jean
Christophe Ianotto, Christophe Marzac, et al. 2016. “Presence of Atypical Thrombopoietin
Receptor (MPL) Mutations in Triple-Negative Essential Thrombocythemia Patients.” Blood 127
(3): 333–42. doi: 10.1182/blood-2015-07-661983.
Cancer Genome Atlas Research Network. Ley, Timothy J., Christopher Miller, Li Ding, Benjamin
J. Raphael, Andrew J. Mungall, A. Gordon Robertson, et al. 2013. “Genomic and Epigenomic
Landscapes of Adult De Novo Acute Myeloid Leukemia.” The New England Journal of Medicine
368 (22): 2059–74. doi: 10.1056/NEJMoa1301689.
Cario, Holger, Jan M. Klaus Schwarz, Vladimir Komrska Herter, Mary F. McMullin, Milen Minkov,
Charlotte Niemeyer, et al. 2008. “Clinical and Molecular Characterisation of a Prospectively
Collected Cohort of Children and Adolescents with Polycythemia Vera.” British Journal of
Haematology 142 (4): 622–6. doi: 10.1111/j.1365-2141.2008.07220.x.
Carron, Clémence, Françoise Cormier, Anne Janin, Virginie Lacronique, Marco Giovannini, Marie-
Thérèse Daniel, Olivier Bernard, and Jacques Ghysdael. 2000. “TEL-JAK2 Transgenic Mice
Develop T-Cell Leukemia.” Blood 95 (12): 3891–9. www.bloodjournal.org/content/95/12/3891.
Cervantes, Francisco, Alberto Alvarez-Larrán, Carme Talarn, Marta Gómez, and Emilio Montserrat.
2002. “Myelofibrosis with Myeloid Metaplasia following Essential Thrombocythaemia: Actuarial
Probability, Presenting Characteristics and Evolution in a Series of 195 Patients.” British Journal of
Haematology 118 (3): 786–90.
Cervantes, Francisco, and Arturo Pereira. 2017. “Does Ruxolitinib Prolong the Survival of Patients with
Myelofibrosis?” Blood 129 (7): 832–7. doi: 10.1182/blood-2016-11-731604.
Chachoua, Ilyas, Christian Pecquet, Mira El-Khoury, Harini Nivarthi, Roxana-Irina Albu,
Caroline Marty, Vitalina Gryshkova, et al. 2016. “Thrombopoietin Receptor Activation by Myelo­
proliferative Neoplasm Associated Calreticulin Mutants.” Blood 127 (10): 1325–35. doi: 10.1182/
blood-2015-11-681932.
Chaligné, R., C. Tonetti, R. Besancenot, L. Roy, C. Marty, P. Mossuz, J.-J. Kiladjian, et al. 2008. “New
Mutations of MPL in Primitive Myelofibrosis: Only the MPL W515 Mutations Promote a
G1/S-Phase Transition.” Leukemia 22 (8): 1557–66. doi: 10.1038/leu.2008.137.
Chaligné, Ronan, Chloé James, Carole Tonetti, Rodolphe Besancenot, Jean Pierre Le Couédic,
Fanny Fava, Fréderic Mazurier, et al. 2007. “Evidence for MPL W515L/K Mutations in Hemato­
poietic Stem Cells in Primitive Myelofibrosis.” Blood 110 (10): 3735–43. doi: 10.1182/blood-2007­
05-089003.
Chapelle, A. de la, A. L. Träskelin, and E. Juvonen. 1993. “Truncated Erythropoietin Receptor Causes
Dominantly Inherited Benign Human Erythrocytosis.” Proceedings of the National Academy of
Sciences of the United States of America 90 (10): 4495–9.
Chapiro, E., L. Russell, E. Lainey, S. Kaltenbach, C. Ragu, V. Della-Valle, K. Hanssens, et al. 2010.
“Activating Mutation in the TSLPR Gene in B-Cell Precursor Lymphoblastic Leukemia.” Leuke­
mia 24 (3): 642–5. doi: 10.1038/leu.2009.231.
Chen, Edwin, Philip A. Beer, Anna L. Godfrey, Christina A. Ortmann, Juan Li, Ana P. Costa-Pereira,
Catherine E. Ingle, Emmanouil T. Dermitzakis, Peter J. Campbell, and Anthony R. Green, 2010.
“Distinct Clinical Phenotypes Associated with JAK2V617F Reflect Differential STAT1 Signaling.”
Cancer Cell 18 (5): 524–35. doi: 10.1016/j.ccr.2010.10.013.
Chen, Jing, Yong Zhang, N. Michael, Wenming Petrus, Alina Xiao, Mark Nicolae, Stefania
Pittaluga Raffeld, et al. 2017. “Cytokine Receptor Signaling Is Required for the Survival of ALK−
Anaplastic Large Cell Lymphoma, Even in the Presence of JAK1/STAT3 Mutations.” Proceedings
of the National Academy of Sciences 114 (15): 3975–80. doi: 10.1073/pnas.1700682114.
Cheng, Ying, Kudakwashe Chikwava, Chao Wu, Haibing Zhang, Anchit Bhagat, Dehua Pei, John
K. Choi, and Wei Tong. 2016. “LNK/SH2B3 Regulates IL-7 Receptor Signaling in Normal and
Malignant B-Progenitors.” The Journal of Clinical Investigation 126 (4): 1267–81. doi: 10.1172/
JCI81468.
Chessells, J. M., G. Harrison, S. M. Richards, C. C. Bailey, F. G. H. Hill, B. E. Gibson, and I. M. Hann.
2001. “Down’s Syndrome and Acute Lymphoblastic Leukaemia: Clinical Features and Response to
Treatment.” Archives of Disease in Childhood 85 (4): 321–5. doi: 10.1136/adc.85.4.321.
Chim, Chor-Sang, Yok-Lam Kwong, Albert Kwok-Wei Lie, Siu-Kwan Ma, Chi-Chung Chan, Lap-Gate
Wong, Bonnie Chi San Kho, et al. 2005. “Long-Term Outcome of 231 Patients with Essential
204 JAK-STAT Signaling in Diseases

Thrombocythemia: Prognostic Factors for Thrombosis, Bleeding, Myelofibrosis, and Leukemia.”


Archives of Internal Medicine 165 (22): 2651–8. doi: 10.1001/archinte.165.22.2651.
Compagno, Mara, Wei Keat Lim, Subhadra V. Adina Grunn, Manisha Brahmachary Nandula,
Qiong Shen, Francesco Bertoni, et al. 2009. “Mutations of Multiple Genes Cause Deregulation of
NF-ΚB in Diffuse Large B-Cell Lymphoma.” Nature 459 (7247): 717–21. doi: 10.1038/nature07968.
Cooper, Todd M., Robert B. Janet Franklin, Todd A. Gerbing, Craig Hurwitz Alonzo, Susana
C. Raimondi, Betsy Hirsch, et al. 2012. “AAML03P1, A Pilot Study of the Safety of Gemtuzumab
Ozogamicin in Combination with Chemotherapy for Newly Diagnosed Childhood Acute Myeloid
Leukemia: A Report from the Children’s Oncology Group.” Cancer 118 (3): 761–9. doi: 10.1002/
cncr.26190.
Cornejo, Melanie G., Michael G. Kharas, Miriam B. Werneck, Séverine Le Bras, Sandra A. Moore,
Brian Ball, Marie Beylot-Barry, et al. 2009. “Constitutive JAK3 Activation Induces Lymphoproli­
ferative Syndromes in Murine Bone Marrow Transplantation Models.” Blood 113 (12): 2746–54.
doi: 10.1182/blood-2008-06-164368.
Cortelazzo, S., P. Viero, G. Finazzi, A. D’Emilio, F. Rodeghiero, and T. Barbui. 1990. “Incidence and
Risk Factors for Thrombotic Complications in a Historical Cohort of 100 Patients with Essential
Thrombocythemia.” Journal of Clinical Oncology: Official Journal of the American Society of
Clinical Oncology 8 (3): 556–62. doi: 10.1200/JCO.1990.8.3.556.
Coskun, Mehmet Enes, Sue Height, Anil Dhawan, and Nedim Hadzic. 2017. “Ruxolitinib Treatment in
an Infant with JAK2+ Polycythaemia Vera-Associated Budd-Chiari Syndrome.” BMJ Case
Reports 2017 (July): bcr-2017-220377. doi: 10.1136/bcr-2017-220377.
Couronné, Lucile, Laurianne Scourzic, Camilla Pilati, Véronique Della Valle, Yannis Duffourd,
Eric Solary, William Vainchenker, et al. 2013. “STAT3 Mutations Identified in Human Hematolo­
gic Neoplasms Induce Myeloid Malignancies in a Mouse Bone Marrow Transplantation Model.”
Haematologica 98 (11): 1748–52. doi: 10.3324/haematol.2013.085068.
Crescenzo, Ramona, Francesco Abate, Elena Lasorsa, Fabrizio Tabbo’, Marcello Gaudiano,
Nicoletta Chiesa, Filomena Di Giacomo, et al. 2015. “Convergent Mutations and Kinase Fusions
Lead to Oncogenic STAT3 Activation in Anaplastic Large Cell Lymphoma.” Cancer Cell 27 (4):
516–32. doi: 10.1016/j.ccell.2015.03.006.
Creutzig, Ursula, Marry M. van Den Heuvel-Eibrink, Michael N. Brenda Gibson, Souichi
Adachi Dworzak, Eveline de Bont, Jochen Harbott, et al. 2012. “Diagnosis and Management of
Acute Myeloid Leukemia in Children and Adolescents: Recommendations from an International
Expert Panel.” Blood 120 (16): 3187–205. doi: 10.1182/blood-2012-03-362608.
Curran, Shane A., Justin A. Shyer, Erin T. St Angelo, Lillian R. Talbot, David J. Sneh Sharma, Glenn
Heller Chung, Katharine C. Hsu, Brian C. Betts, and James W. Young. 2017. “Human Dendritic
Cells Mitigate NK-Cell Dysfunction Mediated by Nonselective JAK1/2 Blockade.” Cancer Immu­
nology Research 5 (1): 52–60. doi: 10.1158/2326-6066.CIR-16-0233.
Debureaux, P. E., J. Arrondeau, D. Bouscary, and F. Goldwasser. 2018. “Nivolumab Combined with
Ruxolitinib: Antagonism or Synergy?” Annals of Oncology 29 (5): 1334–5. doi: 10.1093/annonc/
mdy077.
Degryse, Sandrine, Charles E. de Bock, Luk Cox, Sofie Demeyer, Olga Gielen, Nicole Mentens,
Kris Jacobs, et al. 2014. “JAK3 Mutants Transform Hematopoietic Cells through JAK1 Activation,
Causing T-Cell Acute Lymphoblastic Leukemia in a Mouse Model.” Blood 124 (20): 3092–100. doi:
10.1182/blood-2014-04-566687.
DeLario, Melissa R., Andrea Sheehan, Ramona Ataya, Alison A. Bertuch, Carlos Vega, C. Renee Webb,
Dolores lopez-Terrada, and Lakshmi Venkateswaran. 2010. “Primary Myelofibrosis in Children.”
Blood 116 (21): 3079. www.bloodjournal.org/content/116/21/3079.
DeLario, Melissa R., Andrea M. Sheehan, Ramona Ataya, Alison A. Bertuch, Carlos Vega,
C. Renee Webb, Dolores Lopez-Terrada, and Lakshmi Venkateswaran. 2012. “Clinical, Histo­
pathologic, and Genetic Features of Pediatric Primary Myelofibrosis—An Entity Different from
Adults.” American Journal of Hematology 87 (5): 461–4. doi: 10.1002/ajh.23140.
Den Boer, Monique L., Marjon van Slegtenhorst, Renée X. De Menezes, Meyling H. Cheok, Jessica
G. C. A. M. Buijs-Gladdines, Susan T. C. J. M. Peters, Laura J. C. M. Van Zutven, et al. 2009.
“A Subtype of Childhood Acute Lymphoblastic Leukaemia with Poor Treatment Outcome:
JAK-STAT Signaling in Hematologic Malignancies 205

A Genome-Wide Classification Study.” The Lancet Oncology 10 (2): 125–34. doi: 10.1016/S1470­
2045(08)70339-5.
Derenzini, E., M. Lemoine, D. Buglio, H. Katayama, Y. Ji, R. E. Davis, S. Sen, and A. Younes, 2011.
“The JAK Inhibitor AZD1480 Regulates Proliferation and Immunity in Hodgkin Lymphoma.”
Blood Cancer Journal 1 (12): e46. doi: 10.1038/bcj.2011.46.
Diaz, Tania, Alfons Navarro, Gerardo Ferrer, Bernat Gel, Anna Gaya, Rosa Artells, Beatriz Bellosillo,
et al. 2011. “Lestaurtinib Inhibition of the JAK/STAT Signaling Pathway in Hodgkin Lymphoma
Inhibits Proliferation and Induces Apoptosis.” PLoS One 6 (4): e18856. doi: 10.1371/journal.
pone.0018856.
Ding, Jianmin, Hirokazu Komatsu, Atsushi Wakita, Miyuki Kato-Uranishi, Masato Ito, Atsushi Satoh,
Kazuya Tsuboi, et al. 2004. “Familial Essential Thrombocythemia Associated with a
Dominant-Positive Activating Mutation of the c-MPL Gene, Which Encodes for the Receptor for
Thrombopoietin.” Blood 103 (11): 4198–200. doi: 10.1182/blood-2003-10-3471.
Dördelmann, M., M. Schrappe, A. Reiter, M. Zimmermann, N. Graf, G. Schott, F. Lampert, et al. 1998.
“Down’s Syndrome in Childhood Acute Lymphoblastic Leukemia: Clinical Characteristics and
Treatment Outcome in Four Consecutive BFM Trials.” Leukemia 12 (5): 645. doi: 10.1038/sj.
leu.2400989.
Drachman, J. G., J. D. Griffin, and K. Kaushansky. 1995. “The C-Mpl Ligand (Thrombopoietin)
Stimulates Tyrosine Phosphorylation of Jak2, Shc, and c-Mpl.” The Journal of Biological Chemistry
270 (10): 4979–82.
Dubois, Sydney, Pierre-Julien Viailly, Sylvain Mareschal, Elodie Bohers, Philippe Bertrand,
Philippe Ruminy, Catherine Maingonnat, et al. 2016. “Next-Generation Sequencing in Diffuse
Large B-Cell Lymphoma Highlights Molecular Divergence and Therapeutic Opportunities:
A LYSA Study.” Clinical Cancer Research 22 (12): 2919–28. doi: 10.1158/1078-0432.CCR-15-2305.
Duek, Adrian, Pontus Lundberg, Takafumi Shimizu, Jean Grisouard, Axel Karow, Lucia Kubovcakova,
Hui Hao-Shen, Stephan Dirnhofer, and Radek C. Skoda. 2014. “Loss of Stat1 Decreases Mega­
karyopoiesis and Favors Erythropoiesis in a JAK2-V617F–Driven Mouse Model of MPNs.” Blood
123 (25): 3943–50. doi: 10.1182/blood-2013-07-514208.
Dupont, Sabrina, Aline Massé, Chloé James, Irène Teyssandier, Yann Lécluse, Frédéric Larbret,
Valérie Ugo, et al. 2007. “The JAK2 617V>F Mutation Triggers Erythropoietin Hypersensitivity
and Terminal Erythroid Amplification in Primary Cells from Patients with Polycythemia Vera.”
Blood 110 (3): 1013–21. doi: 10.1182/blood-2006-10-054940.
Eggleton, Paul, and Marek Michalak. 2013. “Calreticulin for Better or for Worse, in Sickness and in
Health, until Death Do Us Part.” Cell Calcium 54 (2): 126–31. doi: 10.1016/j.ceca.2013.05.006.
Eghtedar, Alireza, Srdan Verstovsek, Zeev Estrov, Jan Burger, Jorge Cortes, Carol Bivins, Stefan Faderl,
et al. 2012. “Phase 2 Study of the JAK Kinase Inhibitor Ruxolitinib in Patients with Refractory
Leukemias, Including Postmyeloproliferative Neoplasm Acute Myeloid Leukemia.” Blood 119 (20):
4614–18. doi: 10.1182/blood-2011-12-400051.
Englund, Annika, Ingrid Glimelius, Karin E. Klaus Rostgaard, Sandra Eloranta Smedby, Daniel Molin,
Thomas Kuusk, et al. 2018. “Hodgkin Lymphoma in Children, Adolescents and Young Adults –
A Comparative Study of Clinical Presentation and Treatment Outcome.” Acta Oncologica (Stock­
holm, Sweden) 57 (2): 276–82. doi: 10.1080/0284186X.2017.1355563.
Davis, R. Eric, Keith D. Brown, Ulrich Siebenlist, and Louis M. Staudt. 2001. “Constitutive Nuclear
Factor ΚB Activity Is Required for Survival of Activated B Cell–Like Diffuse Large B Cell
Lymphoma Cells.” Journal of Experimental Medicine 194 (12): 1861–74. doi: 10.1084/
jem.194.12.1861.
Farruggia, Piero, Paolo D’Angelo, Maria La Rosa, Nunzia Scibetta, Giuseppe Santangelo, Antonio
Lo Bello, Elena Duner, Maria Luigia Randi, Maria Caterina Putti, and Alessandra Santoro. 2013.
“MPL W515L Mutation in Pediatric Essential Thrombocythemia.” Pediatric Blood & Cancer 60
(8): E52–4. doi: 10.1002/pbc.24500.
Feenstra, Jelena D., Harini Milosevic, Heinz Gisslinger Nivarthi, Emilie Leroy, Elisa Rumi,
Ilyas Chachoua, Klaudia Bagienski, et al. 2016. “Whole-Exome Sequencing Identifies Novel MPL
and JAK2 Mutations in Triple-Negative Myeloproliferative Neoplasms.” Blood 127 (3): 325–32.
doi: 10.1182/blood-2015-07-661835.
206 JAK-STAT Signaling in Diseases

Fenaux, P., M. Simon, M. T. Caulier, J. L. Lai, J. Goudemand, and F. Bauters. 1990. “Clinical Course of
Essential Thrombocythemia in 147 Cases.” Cancer 66 (3): 549–56.
Flex, Elisabetta, Valentina Petrangeli, Lorenzo Stella, Sabina Chiaretti, Tekla Hornakova,
Laurent Knoops, Cristina Ariola, et al. 2008. “Somatically Acquired JAK1 Mutations in Adult
Acute Lymphoblastic Leukemia.” Journal of Experimental Medicine 205 (4): 751–8. doi: 10.1084/
jem.20072182.
Forestier, Erik, Shai Izraeli, Berna Beverloo, Oskar Haas, Andrea Pession, Kyra Michalová, Batia Stark,
Christine J. Harrison, Andrea Teigler-Schlegel, and Bertil Johansson. 2008. “Cytogenetic Features
of Acute Lymphoblastic and Myeloid Leukemias in Pediatric Patients with Down Syndrome: An
IBFM-SG Study.” Blood 111 (3): 1575–83. doi: 10.1182/blood-2007-09-114231.
Fu, R., D. Liu, Z. Cao, S. Zhu, H. Li, H. Su, L. Zhang, et al. 2016. “Distinct Molecular Abnormalities
Underlie Unique Clinical Features of Essential Thrombocythemia in Children.” Leukemia 30 (3):
746–9. doi: 10.1038/leu.2015.167.
Fu, Rongfeng, Lei Zhang, and Renchi Yang. 2013. “Paediatric Essential Thrombocythaemia: Clinical and
Molecular Features, Diagnosis and Treatment.” British Journal of Haematology 163 (3): 295–302. doi:
10.1111/bjh.12530.
Gaikwad, Amos, Cassia L. Rye, Meenakshi Devidas, Nyla A. Heerema, Andrew J. Carroll, Shai Izraeli,
Sharon E. Plon, Giuseppe Basso, Andrea Pession, and Karen R. Rabin. 2009. “Prevalence and
Clinical Correlates of JAK2 Mutations in Down Syndrome Acute Lymphoblastic Leukaemia.”
British Journal of Haematology 144 (6): 930–2. doi: 10.1111/j.1365-2141.2008.07552.x.
Gamis, Alan S., Todd A. Alonzo, Soheil Meshinchi, Lillian Sung, Robert B. Gerbing, Susana
C. Raimondi, Betsy A. Hirsch, et al. 2014. “Gemtuzumab Ozogamicin in Children and Adolescents
with De Novo Acute Myeloid Leukemia Improves Event-Free Survival by Reducing Relapse Risk:
Results from the Randomized Phase III Children’s Oncology Group Trial AAML0531.” Journal of
Clinical Oncology: Official Journal of the American Society of Clinical Oncology 32 (27): 3021–32.
doi: 10.1200/JCO.2014.55.3628.
Gamis, Alan S., Todd A. Alonzo, John P. Perentesis, and Soheil Meshinchi. 2013. “Children’s Oncology
Group’s 2013 Blueprint for Research: Acute Myeloid Leukemia.” Pediatric Blood & Cancer 60 (6):
964–71. doi: 10.1002/pbc.24432.
Garçon, Loïc, Christine Rivat, Chloé James, Catherine Lacout, Valérie Camara-Clayette, Valérie Ugo,
Yann Lecluse, Annelise Bennaceur-Griscelli, and William Vainchenker. 2006. “Constitutive Activa­
tion of STAT5 and Bcl-XL Overexpression Can Induce Endogenous Erythroid Colony Formation
in Human Primary Cells.” Blood 108 (5): 1551–4. doi: 10.1182/blood-2005-10-009514.
Giona, Fiorina, Luciana Teofili, Sara Capodimonti, Marica Laurino, Maurizio Martini,
Deborah Marzella, Giovanna Palumbo, Daniela Diverio, Robin Foà, and Luigi Maria Larocca.
2014. “CALR Mutations in Patients with Essential Thrombocythemia Diagnosed in Childhood
and Adolescence.” Blood 123 (23): 3677–9. doi: 10.1182/blood-2014-04-572040.
Giona, Fiorina, Luciana Teofili, Maria Luisa Moleti, Maurizio Martini, Giovanna Palumbo,
Angela Amendola, Maria Gabriella Mazzucconi, et al. 2012. “Thrombocythemia and Polycythemia
in Patients Younger than 20 Years at Diagnosis: Clinical and Biologic Features, Treatment, and
Long-Term Outcome.” Blood 119 (10): 2219–27. doi: 10.1182/blood-2011-08-371328.
Goto, Hiroaki, Takeshi Inukai, Hiroyasu Inoue, Chitose Ogawa, Takashi Fukushima, Miharu Yabe,
Akira Kikuchi, et al. 2011. “Acute Lymphoblastic Leukemia and Down Syndrome: The Collabora­
tive Study of the Tokyo Children’s Cancer Study Group and the Kyushu Yamaguchi Children’s
Cancer Study Group.” International Journal of Hematology 93 (2): 192–8. doi: 10.1007/s12185-011­
0765-3.
Green, Michael R., Stefano Monti, Scott J. Rodig, Przemyslaw Juszczynski, Treeve Currie,
Evan O’Donnell, Bjoern Chapuy, et al. 2010. “Integrative Analysis Reveals Selective 9p24.1
Amplification, Increased PD-1 Ligand Expression, and Further Induction via JAK2 in Nodular
Sclerosing Hodgkin Lymphoma and Primary Mediastinal Large B-Cell Lymphoma.” Blood 116
(17): 3268–77. doi: 10.1182/blood-2010-05-282780.
Groner, Bernd, and Viktoria von Manstein. 2017. “Jak Stat Signaling and Cancer: Opportunities, Benefits
and Side Effects of Targeted Inhibition.” Molecular and Cellular Endocrinology, JAK–Stat Signaling
and Cancer 451 (August): 1–14. doi: 10.1016/j.mce.2017.05.033.
JAK-STAT Signaling in Hematologic Malignancies 207

Guglielmelli, P., T. L. Lasho, G. Rotunno, J. Score, C. Mannarelli, A. Pancrazzi, F. Biamonte, et al. 2014.
“The Number of Prognostically Detrimental Mutations and Prognosis in Primary Myelofibrosis:
An International Study of 797 Patients.” Leukemia 28 (9): 1804–10. doi: 10.1038/leu.2014.76.
Gunawan, Arief, Patrick Harrington, Natalia Garcia-Curto, Donal McLornan, Deepti Radia, and
Claire Harrison. 2018. “Ruxolitinib for the Treatment of Essential Thrombocythemia.” Hema-
Sphere 2 (4): e56. doi: 10.1097/HS9.0000000000000056.
Gunawardana, Jay, Fong Chun Chan, Adèle Telenius, Bruce Woolcock, Robert Kridel, King L. Tan,
Susana Ben-Neriah, et al. 2014. “Recurrent Somatic Mutations of PTPN1 in PR and Hodgkin
Lymphoma.” Nature Genetics 46 (4): 329–35. doi: 10.1038/ng.2900.
Gupta, Mamta, Jing Jing Han, Mary Stenson, Matthew Maurer, Linda Wellik, Guangzhen Hu,
Steve Ziesmer, Ahmet Dogan, and Thomas E. Witzig. 2012. “Elevated Serum IL-10 Levels in
Diffuse Large B-Cell Lymphoma: A Mechanism of Aberrant JAK2 Activation.” Blood 119 (12):
2844–53. doi: 10.1182/blood-2011-10-388538.
Hama, Asahito, Hiroshi Yagasaki, Yoshiyuki Takahashi, Kimikazu Matsumoto, Hitoshi Kiyoi, and
Seiji Kojima. 2008. “Response: Mutations of JAK2, JAK3 and GATA1 in Acute Megakaryoblastic
Leukemia of down Syndrome.” Blood 111 (4): 2493–4. doi: 10.1182/blood-2007-10-117614.
Hamlin, Paul, Bruce Cheson, Charles Farber, Tatyana Feldman, Timothy Fenske, Greg Coffey, and
John Curnutte. 2018. “The Dual SYK/JAK Inhibitor Cerdulatinib Demonstrates Rapid Tumor
Responses in a Phase 2 Study in Patients with Relapsed/Refractory B- and T-Cell Non-Hodgkin
Lymphoma (NHL).” Journal of Clinical Oncology ASCO 2018 Abstract.
Hao, Yansheng, Bjoern Chapuy, Stefano Monti, Heather H. Sun, Scott J. Rodig, and Margaret A. Ship.
2014. “Selective JAK2 Inhibition Specifically Decreases Hodgkin Lymphoma and Mediastinal
Large B-Cell Lymphoma Growth in Vitro and in Vivo.” Clinical Cancer Research 20 (10): 2674–83.
doi: 10.1158/1078-0432.CCR-13-3007.
Harrison, Claire N., Adam J. Mead, Anesh Panchal, Sonia Fox, Christina Yap, Emmanouela Gbandi,
Aimee Houlton, et al. 2017a. “Ruxolitinib vs Best Available Therapy for ET Intolerant or Resistant
to Hydroxycarbamide.” Blood 130 (17): 1889–97. doi: 10.1182/blood-2017-05-785790.
Harrison, Claire N, Nicolaas Schaap, Alessandro M Vannucchi, Jean-Jacques Kiladjian, Ramon V Tiu,
Pierre Zachee, Eric Jourdan, et al. 2017b. “Janus Kinase-2 Inhibitor Fedratinib in Patients with
Myelofibrosis Previously Treated with Ruxolitinib (JAKARTA-2): A Single-Arm, Open-Label,
Non-Randomised, Phase 2, Multicentre Study.” The Lancet Haematology 4 (7): e317–24. doi:
10.1016/S2352-3026(17)30088-1.
Hart, S., K. C. Goh, V. Novotny-Diermayr, C. Y. Hu, H. Hentze, Y. C. Tan, B. Madan, et al. 2011.
“SB1518, a Novel Macrocyclic Pyrimidine-Based JAK2 Inhibitor for the Treatment of Myeloid and
Lymphoid Malignancies.” Leukemia 25 (11): 1751–9. doi: 10.1038/leu.2011.148.
Harvey, Richard C., Charles G. Mullighan, I.-Ming Chen, Walker Wharton, Fady M. Mikhail, Andrew
J. Carroll, Huining Kang, et al. 2010. “Rearrangement of CRLF2 Is Associated with Mutation of
JAK Kinases, Alteration of IKZF1, Hispanic/Latino Ethnicity, and a Poor Outcome in Pediatric
B-Progenitor Acute Lymphoblastic Leukemia.” Blood 115 (26): 5312–21. doi: 10.1182/blood-2009­
09-245944.
Hasle, H., G. Kerndrup, and B. B. Jacobsen. 1995. “Childhood Myelodysplastic Syndrome in Denmark:
Incidence and Predisposing Conditions.” Leukemia 9 (9): 1569–72.
Hasle, H., L. D. Wadsworth, B. G. Massing, M. McBride, and K. R. Schultz. 1999. “A Population-Based
Study of Childhood Myelodysplastic Syndrome in British Columbia, Canada.” British Journal of
Haematology 106 (4): 1027–32.
Hasle, Henrik. 2000. “Incidence of Essential Thrombocythaemia in Children.” British Journal of
Haematology 110 (3): 751. doi: 10.1046/j.1365-2141.2000.02239-7.x.
Hasle, Henrik, Inge Haunstrup Clemmensen, and Margareta Mikkelsen. 2000. “Risks of Leukaemia and
Solid Tumours in Individuals with Down’s Syndrome.” The Lancet 355 (9199): 165–9. doi: 10.1016/
S0140-6736(99)05264-2.
Hasselbalch, Hans Carl. 2012. “Perspectives on Chronic Inflammation in Essential Thrombocythemia,
Polycythemia Vera, and Myelofibrosis: Is Chronic Inflammation a Trigger and Driver of Clonal
Evolution and Development of Accelerated Atherosclerosis and Second Cancer?” Blood 119 (14):
3219–25. doi: 10.1182/blood-2011-11-394775.
208 JAK-STAT Signaling in Diseases

Hasselbalch, Hans Carl. 2013. “Chronic Inflammation as a Promotor of Mutagenesis in Essential


Thrombocythemia, Polycythemia Vera and Myelofibrosis. A Human Inflammation Model for
Cancer Development?” Leukemia Research 37 (2): 214–20. doi: 10.1016/j.leukres.2012.10.020.
Hennighausen, Lothar, and Gertraud W. Robinson. 2008. “Interpretation of Cytokine Signaling through
the Transcription Factors STAT5A and STAT5B.” Genes & Development 22 (6): 711–21. doi:
10.1101/gad.1643908.
Hertzberg, Libi, Elena Vendramini, Ithamar Ganmore, Gianni Cazzaniga, Maike Schmitz, Jane Chalker,
Ruth Shiloh, et al. 2010. “Down Syndrome Acute Lymphoblastic Leukemia, A Highly Hetero­
geneous Disease in Which Aberrant Expression of CRLF2 Is Associated with Mutated JAK2:
A Report from the International BFM Study Group.” Blood 115 (5): 1006–17. doi: 10.1182/blood­
2009-08-235408.
Hidalgo-López, Juliana E., L. Rashmi Kanagal-Shamanna, Jeffrey Medeiros, C. Zeev Estrov,
Cameron Yin, Srdan Verstovsek, Sergej Konoplev, et al. 2017. “Morphologic and Molecular
Characteristics of De Novo AML with JAK2 V617F Mutation.” Journal of the National Compre­
hensive Cancer Network: JNCCN 15 (6): 790–6. doi: 10.6004/jnccn.2017.0106.
Hornakova, Tekla, Judith Staerk, Yohan Royer, Elisabetta Flex, Stefan N. Marco Tartaglia, Laurent
Knoops Constantinescu, and Jean-Christophe Renauld. 2009. “Acute Lymphoblastic Leukemia-
Associated JAK1 Mutants Activate the Janus Kinase/STAT Pathway via Interleukin-9 Receptor α
Homodimers.” Journal of Biological Chemistry 284 (11): 6773–81. doi: 10.1074/jbc.M807531200.
Hu, Guangzhen, Thomas E. Witzig, and Mamta Gupta. 2013. “A Novel Missense (M206K) STAT3
Mutation in Diffuse Large B Cell Lymphoma Deregulates STAT3 Signaling.” PLoS One 8 (7):
e67851. doi: 10.1371/journal.pone.0067851.
Hunger, Stephen P., and Charles G. Mullighan. 2015a. “Acute Lymphoblastic Leukemia in Children.”
New England Journal of Medicine 373 (16): 1541–52. doi: 10.1056/NEJMra1400972.
Iacobucci, Ilaria, Li Yongjin, Kathryn G. Roberts, Stephanie M. Dobson, Jaeseung C. Kim,
Debbie Payne-Turner, Richard C. Harvey, et al. 2016. “Truncating Erythropoietin Receptor
Rearrangements in Acute Lymphoblastic Leukemia.” Cancer Cell 29 (2): 186–200. doi: 10.1016/j.
ccell.2015.12.013.
Isaksen, Deborah E., Heinz Baumann, Patty A. Trobridge, Andrew G. Farr, Steven D. Levin, and Steven
F. Ziegler. 1999. “Requirement for Stat5 in Thymic Stromal Lymphopoietin-Mediated Signal
Transduction.” The Journal of Immunology 163 (11): 5971–7. www.jimmunol.org/content/163/
11/5971.
Isaksen, Deborah E., Heinz Baumann, Baohua Zhou, Sebastien Nivollet, Andrew G. Farr, Steven
D. Levin, and Steven F. Ziegler. 2002. “Uncoupling of Proliferation and Stat5 Activation in
Thymic Stromal Lymphopoietin-Mediated Signal Transduction.” The Journal of Immunology 168
(7): 3288–94. doi: 10.4049/jimmunol.168.7.3288.
Ishikawa, Chie, Masachika Senba, and Naoki Mori. 2018. “Anti-Adult T-Cell Leukemia/Lymphoma
Activity of Cerdulatinib, a Dual SYK/JAK Kinase Inhibitor.” International Journal of Oncology 53
(4): 1681–90. doi: 10.3892/ijo.2018.4513.
Ismael, Olfat, Akira Shimada, Asahito Hama, Hiroshi Sakaguchi, Sayoko Doisaki, Hideki Muramatsu,
Nao Yoshida, et al. 2012. “Mutations Profile of Polycythemia Vera and Essential Thrombocythe­
mia among Japanese Children.” Pediatric Blood & Cancer 59 (3): 530–5. doi: 10.1002/pbc.23409.
Jain, Nitin, Kathryn G. Roberts, Elias Jabbour, Keyur Patel, Agda Karina Eterovic, Ken Chen,
Patrick Zweidler-McKay, et al. 2017. “Ph-like Acute Lymphoblastic Leukemia: A High-Risk
Subtype in Adults.” Blood 129 (5): 572–81. doi: 10.1182/blood-2016-07-726588.
James, Chloé, Valérie Ugo, Jean-Pierre Le Couédic, Judith Staerk, François Delhommeau,
Catherine Lacout, Loïc Garçon, et al. 2005. “A Unique Clonal JAK2 Mutation Leading to
Constitutive Signalling Causes Polycythaemia Vera.” Nature 434 (7037): 1144–8. doi: 10.1038/
nature03546.
Jamieson, Catriona H. M., Jeffrey A. Jason Gotlib, Mark P. Durocher, Chao M. Rajan Mariappan,
Marla Lay, Carol Jones, James L. Zehnder, Stan L. Lilleberg, and Irving L. Weissman. 2006. “The
JAK2 V617F Mutation Occurs in Hematopoietic Stem Cells in Polycythemia Vera and Predisposes
toward Erythroid Differentiation.” Proceedings of the National Academy of Sciences 103 (16): 6224–9.
doi: 10.1073/pnas.0601462103.
JAK-STAT Signaling in Hematologic Malignancies 209

Jelinek, Jaroslav, Yasuhiro Oki, Vazganush Gharibyan, Josef T. Carlos Bueso-Ramos, Srdan
Verstovsek Prchal, Miloslav Beran, Elihu Estey, Hagop M. Kantarjian, and Jean-Pierre J. Issa.
2005. “JAK2 Mutation 1849G>T Is Rare in Acute Leukemias but Can Be Found in CMML,
Philadelphia Chromosome–Negative CML, and Megakaryocytic Leukemia.” Blood 106 (10): 3370–3.
doi: 10.1182/blood-2005-05-1800.
Jensen, M. K., P. de Nully Brown, O. J. Nielsen, and H. C. Hasselbalch. 2000. “Incidence, Clinical
Features and Outcome of Essential Thrombocythaemia in a Well Defined Geographical Area.”
European Journal of Haematology 65 (2): 132–9.
Jones, Amy V., Peter J. Campbell, Philip A. Beer, Alessandro M. Susanne Schnittger, Katerina
Zoi Vannucchi, MelanieJ. Percy, et al. 2010. “The JAK2 46/1 Haplotype Predisposes to
MPL-Mutated Myeloproliferative Neoplasms.” Blood 115 (22): 4517–23. doi: 10.1182/blood-2009­
08-236448.
Ju, Wei, Kelli M. Meili Zhang, Michael N. Wilson, Richard N. Petrus, Xiaohu Zhang Bamford,
Rajarshi Guha, Marc Ferrer, Craig J. Thomas, and Thomas A. Waldmann. 2016. “Augmented
Efficacy of Brentuximab Vedotin Combined with Ruxolitinib and/or Navitoclax in a Murine Model
of Human Hodgkin’s Lymphoma.” Proceedings of the National Academy of Sciences 113 (6): 1624–9.
doi: 10.1073/pnas.1524668113.
Karjalainen, Riikka, Tea Pemovska, Mihaela Popa, Komal K. Minxia Liu, Muntasir M. Javarappa,
Bhagwan Yadav Majumder, et al. 2017. “JAK1/2 and BCL2 Inhibitors Synergize to Counteract
Bone Marrow Stromal Cell–Induced Protection of AML.” Blood 130 (6): 789–802. doi: 10.1182/
blood-2016-02-699363.
Kawamura, Machiko, Tomohiko Taki, Hidefumi Kaku, Kentaro Ohki, and Yasuhide Hayashi. 2015.
“Identification of SPAG9 as a Novel JAK2 Fusion Partner Gene in Pediatric Acute Lymphoblastic
Leukemia with t(9;17)(P24;Q21).” Genes, Chromosomes and Cancer 54 (7): 401–8. doi: 10.1002/
gcc.22251.
Kearney, Lyndal, David Gonzalez De Castro, Jenny Yeung, Julia Procter, Sharon W. Horsley,
Minenori Eguchi-Ishimae, Caroline M. Bateman, et al. 2009. “Specific JAK2 Mutation
(JAK2R683) and Multiple Gene Deletions in Down Syndrome Acute Lymphoblastic Leukemia.”
Blood 113 (3): 646–8. doi: 10.1182/blood-2008-08-170928.
Kelly, Kevin, Corrina McMahon, Stephen Langabeer, Hassan Eliwan, Aengus O’Marcaigh, and Owen
P. Smith. 2008. “Congenital JAK2V617F Polycythemia Vera: Where Does the Genotype-Phenotype
Diversity End?” Blood 112 (10): 4356–7. doi: 10.1182/blood-2008-08-175620.
Keohane, Clodagh, Shahram Kordasti, Thomas Seidl, Pilar Perez, Nicholas Abellan, S. B. Thomas,
Claire N. Harrison, Donal P. McLornan, and Ghulam J. Mufti. 2015. “JAK Inhibition Induces
Silencing of T Helper Cytokine Secretion and a Profound Reduction in T Regulatory Cells.” British
Journal of Haematology 171 (1): 60–73. doi: 10.1111/bjh.13519.
Khan, I., Z. Huang, Q. Wen, M. J. Stankiewicz, L. Gilles, B. Goldenson, R. Schultz, et al. 2013. “AKT Is
a Therapeutic Target in Myeloproliferative Neoplasms.” Leukemia 27 (9): 1882–90. doi: 10.1038/
leu.2013.167.
Khwaja, Asim. 2006. “The Role of Janus Kinases in Haemopoiesis and Haematological Malignancy.”
British Journal of Haematology 134 (4): 366–84. doi: 10.1111/j.1365-2141.2006.06206.x.
Kim, Sang-Kyu, Deborah A. Knight, Lisa R. Jones, Ashley P. Stephin Vervoort, John F. Ng, James
E. Seymour, Michaela Waibel Bradner, Lev Kats, and Ricky W. Johnstone. 2018. “JAK2 Is
Dispensable for Maintenance of JAK2 Mutant B-Cell Acute Lymphoblastic Leukemias.” Genes &
Development 32 (11–12): 849–64. doi: 10.1101/gad.307504.117.
Kim, Seok Jin, Hye Jin Kang, Dong-Yeop Shin, Dok Hyun Yoon, Kana Sakamoto, Jee Hyun Kong,
Young Hyeh Ko, Kengo Takeuchi, Cheolwon Suh, and Won Seog Kim. 2014. “Pilot Study of
Ruxolitinib in Relapsed or Refractory Hodgkin Lymphoma and Primary Mediastinal Large B-Cell
Lymphoma.” Blood 124 (21): 4443 ASH 2014 Abstract. www.bloodjournal.org/content/124/21/
4443.
Kiyoi, H., S. Yamaji, S. Kojima, and T. Naoe. 2007. “JAK3 Mutations Occur in Acute Megakaryoblastic
Leukemia Both in Down Syndrome Children and Non-Down Syndrome Adults.” Leukemia 21 (3):
574–6. doi: 10.1038/sj.leu.2404527.
Klampfl, Thorsten, Ashot Harutyunyan, Tiina Berg, Bettina Gisslinger, Martin Schalling,
Klaudia Bagienski, Damla Olcaydu, et al. 2011. “Genome Integrity of Myeloproliferative Neoplasms
210 JAK-STAT Signaling in Diseases

in Chronic Phase and during Disease Progression.” Blood 118 (1): 167–76. doi: 10.1182/blood-2011­
01-331678.
Klampfl, Thorsten, Ashot S. Heinz Gisslinger, Harini Nivarthi Harutyunyan, Jelena D. Elisa Rumi,
Nicole C. C. Them Milosevic, et al. 2013. “Somatic Mutations of Calreticulin in Myeloproliferative
Neoplasms.” The New England Journal of Medicine 369 (25): 2379–90. doi: 10.1056/
NEJMoa1311347.
Kleppe, Maria, Thomas Tousseyn, Eva Geissinger, Zeynep Kalender Atak, Stein Aerts,
Andreas Rosenwald, Iwona Wlodarska, and Jan Cools. 2011. “Mutation Analysis of the Tyrosine
Phosphatase PTPN2 in Hodgkin’s Lymphoma and T-Cell Non-Hodgkin’s Lymphoma.” Haemato­
logica 96 (11): 1723–7. doi: 10.3324/haematol.2011.041921.
Koo, Ghee Chong, Soo Yong Tan, Tiffany Tang, Song Ling Poon, George E. Allen, Leonard Tan, Soo
Ching Chong, et al. 2012. “Janus Kinase 3–Activating Mutations Identified in Natural Killer/T-Cell
Lymphoma.” Cancer Discovery 2 (7): 591–7. doi: 10.1158/2159-8290.CD-12-0028.
Koppikar, Priya, Neha Bhagwat, Outi Kilpivaara, Taghi Manshouri, Mazhar Adli, Todd Hricik, Fan Liu,
et al. 2012. “Heterodimeric JAK–STAT Activation as a Mechanism of Persistence to JAK2
Inhibitor Therapy.” Nature 489 (7414): 155–9. doi: 10.1038/nature11303.
Koskela, Hanna L.M., Samuli Eldfors, Arjan J. Pekka Ellonen, Heikki van Adrichem, Emma
I. Kuusanmäki, Sonja Lagström Andersson, et al. 2012. “Somatic STAT3 Mutations in Large
Granular Lymphocytic Leukemia.” New England Journal of Medicine 366 (20): 1905–13. doi:
10.1056/NEJMoa1114885.
Kralovics, Robert, Yongli Guan, and Josef T Prchal. 2002. “Acquired Uniparental Disomy of Chromo­
some 9p Is a Frequent Stem Cell Defect in Polycythemia Vera.” Experimental Hematology 30 (3):
229–36. doi: 10.1016/S0301-472X(01)00789-5.
Kralovics, Robert, Francesco Passamonti, S. Buser Andreas, Soon-Siong Teo, Ralph Tiedt, Jakob
R. Passweg, Andre Tichelli, Mario Cazzola, and Radek C. Skoda. 2005. “A Gain-of-Function
Mutation of JAK2 in Myeloproliferative Disorders.” The New England Journal of Medicine 352
(17): 1779–90. doi: 10.1056/NEJMoa051113.
Kralovics, Robert, David W. Stockton, and Josef T. Prchal. 2003. “Clonal Hematopoiesis in Familial
Polycythemia Vera Suggests the Involvement of Multiple Mutational Events in the Early Pathogen­
esis of the Disease.” Blood 102 (10): 3793–6. doi: 10.1182/blood-2003-03-0885.
Lacout, Catherine, Didier F. Pisani, Micheline Tulliez, Françoise Moreau Gachelin,
William Vainchenker, and Jean-Luc Villeval. 2006. “JAK2V617F Expression in Murine Hemato­
poietic Cells Leads to MPD Mimicking Human PV with Secondary Myelofibrosis.” Blood 108 (5):
1652–60. doi: 10.1182/blood-2006-02-002030.
Lacronique, Virginie, Anthony Boureux, Véronique Della Valle, Hélène Poirel, Christine Tran Quang,
Martine Mauchauffé, Christian Berthou, et al. 1997. “A TEL-JAK2 Fusion Protein with Constitu­
tive Kinase Activity in Human Leukemia.” Science 278 (5341): 1309–12. doi: 10.1126/
science.278.5341.1309.
Lai, Catherine, and Kieron Dunleavy. 2013. “NK/T-Cell Lymphomas in Children.” Best Practice &
Research. Clinical Haematology 26 (1): 33–41. doi: 10.1016/j.beha.2013.04.004.
Lam, Lloyd T., R. George Wright, Eric Davis, Georg Lenz, Pedro Farinha, John W. Lenny Dang,
Andreas Rosenwald Chan, Randy D. Gascoyne, and Louis M. Staudt. 2008. “Cooperative Signal­
ing through the Signal Transducer and Activator of Transcription 3 and Nuclear Factor-ΚB
Pathways in Subtypes of Diffuse Large B-Cell Lymphoma.” Blood 111 (7): 3701–13. doi: 10.1182/
blood-2007-09-111948.
Lasho, Terra L., Rebecca F. Animesh Pardanani, Ruben A. McClure, Ross L. Mesa, Levine D. Gary
Gilliland, and Ayalew Tefferi. 2006. “Concurrent MPL515 and JAK2V617F Mutations in Myelofi­
brosis: Chronology of Clonal Emergence and Changes in Mutant Allele Burden over Time.” British
Journal of Haematology 135 (5): 683–7. doi: 10.1111/j.1365-2141.2006.06348.x.
Lee, Sanghoon, Tishi Shah, Changhong Yin, Jessica Hochberg, Janet Ayello, Erin Morris, Carmella van
de Ven, and Mitchell S. Cairo. 2018. “Ruxolitinib Significantly Enhances in Vitro Apoptosis in
Hodgkin Lymphoma and Primary Mediastinal B-Cell Lymphoma and Survival in a Lymphoma
Xenograft Murine Model.” Oncotarget 9 (11): 9776–88. doi: 10.18632/oncotarget.24267.
Lenglet, Marion, Florence Robriquet, Klaus Schwarz, Carme Camps, Anne Couturier, David Hoogewijs,
Alexandre Buffet, et al. 2018. “Identification of a New VHL Exon and Complex Splicing
JAK-STAT Signaling in Hematologic Malignancies 211

Alterations in Familial Erythrocytosis or Von Hippel-Lindau Disease.” Blood 132 (5): 469–83. doi:
10.1182/blood-2018-03-838235.
Lenz, Georg, George W. Wright, N. C. Tolga Emre, Sandeep S. Holger Kohlhammer, Dave R. Eric Davis,
Shannon Carty, et al. 2008. “Molecular Subtypes of Diffuse Large B-Cell Lymphoma Arise by
Distinct Genetic Pathways.” Proceedings of the National Academy of Sciences 105 (36): 13520–5.
doi: 10.1073/pnas.0804295105.
Levine, Ross L., Martha Wadleigh, Benjamin L. Jan Cools, Gerlinde Wernig Ebert, Brian J. P. Huntly,
Titus J. Boggon, et al. 2005. “Activating Mutation in the Tyrosine Kinase JAK2 in Polycythemia
Vera, Essential Thrombocythemia, and Myeloid Metaplasia with Myelofibrosis.” Cancer Cell 7 (4):
387–97. doi: 10.1016/j.ccr.2005.03.023.
Lippert, Eric, Marjorie Boissinot, Robert Kralovics, François Girodon, Irène Dobo, Vincent Praloran,
Nathalie Boiret-Dupré, Radek C. Skoda, and Sylvie Hermouet. 2006. “The JAK2-V617F Mutation
Is Frequently Present at Diagnosis in Patients with Essential Thrombocythemia and Polycythemia
Vera.” Blood 108 (6): 1865–7. doi: 10.1182/blood-2006-01-013540.
Liu, Yu, John Easton, Ying Shao, Jamie Maciaszek, Mark R. Zhaoming Wang, Kelly
McCastlain Wilkinson, et al. 2017. “The Genomic Landscape of Pediatric and Young Adult
T-Lineage Acute Lymphoblastic Leukemia.” Nature Genetics 49 (8): 1211–18. doi: 10.1038/ng.3909.
Loh, Mignon L., Sarah K. Tasian, Karen R. Rabin, Patrick Brown, Joel M. Daniel Magoon, Xuejun
Chen Reid, Charlotte H. Ahern, Brenda J. Weigel, and Susan M. Blaney. 2015. “A Phase 1 Dosing
Study of Ruxolitinib in Children with Relapsed or Refractory Solid Tumors, Leukemias, or
Myeloproliferative Neoplasms: A Children’s Oncology Group Phase 1 Consortium Study
(ADVL1011).” Pediatric Blood & Cancer 62 (10): 1717–24. doi: 10.1002/pbc.25575.
Lohr, Jens G., Michael S. Petar Stojanov, Daniel Auclair Lawrence, Bjoern Chapuy, Carrie Sougnez,
Peter Cruz-Gordillo, et al. 2012. “Discovery and Prioritization of Somatic Mutations in Diffuse
Large B-Cell Lymphoma (DLBCL) by Whole-Exome Sequencing.” Proceedings of the National
Academy of Sciences 109 (10): 3879–84. doi: 10.1073/pnas.1121343109.
Lorsbach, Robert B., Dominic Shay-Seymore, Jennifer Moore, Peter M. Banks, Robert P. Hasserjian,
John T. Sandlund, and Frederick G. Behm. 2002. “Clinicopathologic Analysis of Follicular
Lymphoma Occurring in Children.” Blood 99 (6): 1959–64. doi: 10.1182/blood.V99.6.1959.
Louissaint, Abner, Kristian T. Schafernak, Julia T. Geyer, Alexandra E. Kovach, Mahmoud Ghandi,
Dita Gratzinger, Christine G. Roth, et al. 2016. “Pediatric-Type Nodal Follicular Lymphoma:
A Biologically Distinct Lymphoma with Frequent MAPK Pathway Mutations.” Blood 128 (8):
1093–100. doi: 10.1182/blood-2015-12-682591.
Lundberg, Pontus, Axel Karow, Ronny Nienhold, Renate Looser, Hui Hao-Shen, Ina Nissen,
Sabine Girsberger, et al. 2014. “Clonal Evolution and Clinical Correlates of Somatic Mutations in
Myeloproliferative Neoplasms.” Blood 123 (14): 2220–8. doi: 10.1182/blood-2013-11-537167.
Ma, Jiao, Wei Xing, Greg Coffey, Karen Dresser, Lu Kellie, Ailin Guo, Gordana Raca, et al. 2015.
“Cerdulatinib, a Novel Dual SYK/JAK Kinase Inhibitor, Has Broad Anti-Tumor Activity in Both
ABC and GCB Types of Diffuse Large B Cell Lymphoma.” Oncotarget 6 (41): 43881–96. doi:
10.18632/oncotarget.6316.
Malinge, Sebastien, Raouf Ben-Abdelali, Catherine Settegrana, Isabelle Radford-Weiss, Marianne Debre,
Kheira Beldjord, Elizabeth A. Macintyre, et al. 2007. “Novel Activating JAK2 Mutation in
a Patient with Down Syndrome and B-Cell Precursor Acute Lymphoblastic Leukemia.” Blood 109
(5): 2202–4. doi: 10.1182/blood-2006-09-045963.
Malinge, Sébastien, Christine Ragu, Veronique Della-Valle, Stefan N. Didier Pisani,
Christelle Constantinescu, Jean-Luc Villeval Perez, et al. 2008. “Activating Mutations in Human
Acute Megakaryoblastic Leukemia.” Blood 112 (10): 4220–6. doi: 10.1182/blood-2008-01-136366.
Marchioli, Roberto, Guido Finazzi, Raffaele Landolfi, Jack Kutti, Heinz Gisslinger, Carlo Patrono,
Raphael Marilus, Ana Villegas, Gianni Tognoni, and Tiziano Barbui. 2005. “Vascular and Neo­
plastic Risk in a Large Cohort of Patients with Polycythemia Vera.” Journal of Clinical Oncology 23
(10): 2224–32. doi: 10.1200/JCO.2005.07.062.
Marshall, John L., Hedy Kindler, John Deeken, Nicholas J. Pankaj Bhargava, Naiyer Rizvi Vogelzang,
Taina Luhtala, et al. 2005. “Phase I Trial of Orally Administered CEP-701, a Novel Neurotrophin
Receptor-Linked Tyrosine Kinase Inhibitor.” Investigational New Drugs 23 (1): 31–7. doi: 10.1023/
B:DRUG.0000047103.64335.b0.
212 JAK-STAT Signaling in Diseases

Martinaud, Christophe, Patrick Brisou, and Marie-Joelle Mozziconacci. 2010. “Is the JAK2(V617F)
Mutation Detectable in Healthy Volunteers?” American Journal of Hematology 85 (4): 287–8. doi:
10.1002/ajh.21627.
Martínez-Avilés, Luz, Carles Besses, Alberto Álvarez-Larrán, Francisco Cervantes, Juan
Carlos Hernández-Boluda, and Beatriz Bellosillo. 2007. “JAK2 Exon 12 Mutations in Polycythemia
Vera or Idiopathic Erythrocytosis.” Haematologica 92 (12): 1717–18. doi: 10.3324/haematol.12011.
Marty, Caroline, Christian Pecquet, Harini Nivarthi, Mira El-Khoury, Ilyas Chachoua,
Micheline Tulliez, Jean-Luc Villeval, et al. 2016. “Calreticulin Mutants in Mice Induce an
MPL-Dependent Thrombocytosis with Frequent Progression to Myelofibrosis.” Blood 127 (10):
1317–24. doi: 10.1182/blood-2015-11-679571.
Mascarenhas, John, Ronald Hoffman, Aaron T. Moshe Talpaz, Brady Stein Gerds, Vikas Gupta,
Anita Szoke, et al. 2018. “Pacritinib Vs Best Available Therapy, Including Ruxolitinib, in Patients
with Myelofibrosis: A Randomized Clinical Trial.” JAMA Oncology 4 (5): 652–9. doi: 10.1001/
jamaoncol.2017.5818.
Maslah, N., B. Cassinat, E. Verger, J.-J. Kiladjian, and L. Velazquez. 2017. “The Role of LNK/SH2B3
Genetic Alterations in Myeloproliferative Neoplasms and Other Hematological Disorders.” Leu­
kemia 31 (8): 1661–70. doi: 10.1038/leu.2017.139.
Mathas, Stephan, Sylvia Hartmann, and Ralf Küppers. 2016. “Hodgkin Lymphoma: Pathology and
Biology.” Seminars in Hematology, Hodgkin Lymphoma 53 (3): 139–47. doi: 10.1053/j.
seminhematol.2016.05.007.
Maude, Shannon L., Sarah K. Tasian, Junior W. Tiffaney Vincent, Cecilia Sheen Hall, Kathryn G. Roberts,
AlixE Seif, et al. 2012. “Targeting JAK1/2 and MTOR in Murine Xenograft Models of Ph-Like Acute
Lymphoblastic Leukemia.” Blood 120 (17): 3510–18. doi: 10.1182/blood-2012-03-415448.
McGirt, Laura Y., Peilin Jia, Devin A. Baerenwald, Robert J. Duszynski, Kimberly B. Dahlman, John
A. Zic, Jeffrey P. Zwerner, et al. 2015. “Whole-Genome Sequencing Reveals Oncogenic Mutations
in Mycosis Fungoides.” Blood 126 (4): 508–19. doi: 10.1182/blood-2014-11-611194.
McMullin, Mary Frances, and Holger Cario. 2016. “LNK Mutations and Myeloproliferative Disorders.”
American Journal of Hematology 91 (2): 248–51. doi: 10.1002/ajh.24259.
McNally, R. J., D. Rowland, E. Roman, and R. A. Cartwright. 1997. “Age and Sex Distributions of
Hematological Malignancies in the U.K.” Hematological Oncology 15 (4): 173–89.
Mead, Adam J., Michelle J. Rugless, W. Jacobsen Sten Eirik, and Anna Schuh. 2012. “Germline JAK2
Mutation in a Family with Hereditary Thrombocytosis.” The New England Journal of Medicine 366
(10): 967–9. doi: 10.1056/NEJMc1200349.
Meier, Cecile, Sylvia Hoeller, Caroline Bourgau, Petra Hirschmann, Juerg Schwaller, Stefano A. Philip
Went, Andreas Reiter Pileri, Stephan Dirnhofer, and Alexandar Tzankov. 2009. “Recurrent
Numerical Aberrations of JAK2 and Deregulation of the JAK2-STAT Cascade in Lymphomas.”
Modern Pathology 22 (3): 476–87. doi: 10.1038/modpathol.2008.207.
Melzner, I., M. A. Weniger, C. K. Menz, and P. Möller. 2006. “Absence of the JAK2 V617F Activating
Mutation in Classical Hodgkin Lymphoma and Primary Mediastinal B-Cell Lymphoma.” Leuke­
mia 20 (1): 157–8. doi: 10.1038/sj.leu.2404036.
Melzner, Ingo, Alexandra Juliana Bucur, Silke Brüderlein, Karola Dorsch, Thomas F. Cornelia Hasel,
E. Barth, Frank Leithäuser, and Möller. Peter. 2005. “Biallelic Mutation of SOCS-1 Impairs JAK2
Degradation and Sustains Phospho-JAK2 Action in the MedB-1 Mediastinal Lymphoma Line.”
Blood 105 (6): 2535–42. doi: 10.1182/blood-2004-09-3701.
Mesa, Ruben A., John V. Jean-Jacques Kiladjian, Timothy Devos Catalano, Miklos Egyed,
Andrzei Hellmann, Donal McLornan, et al. 2017. “SIMPLIFY-1: A Phase III Randomized Trial
of Momelotinib versus Ruxolitinib in Janus Kinase Inhibitor–Naïve Patients with Myelofibrosis.”
Journal of Clinical Oncology 35 (34): 3844–50. doi: 10.1200/JCO.2017.73.4418.
Morak, Maria, Andishe Attarbaschi, Susanna Fischer, Christine Nassimbeni, Reinhard Grausenburger,
Stephan Bastelberger, Stefanie Krentz, et al. 2012. “Small Sizes and Indolent Evolutionary Dynamics
Challenge the Potential Role of P2RY8-CRLF2–Harboring Clones as Main Relapse-Driving Force in
Childhood ALL.” Blood 120 (26): 5134–42. doi: 10.1182/blood-2012-07-443218.
Morin, Ryan D., Andrew J. Maria Mendez-Lago, Rodrigo Goya Mungall, Karen L. Mungall, Richard
D. Corbett, Nathalie A. Johnson, et al. 2011. “Frequent Mutation of Histone-Modifying Genes in
Non-Hodgkin Lymphoma.” Nature 476 (7360): 298–303. doi: 10.1038/nature10351.
JAK-STAT Signaling in Hematologic Malignancies 213

Mottok, Anja, Christoph Renné, Marc Seifert, Elsie Oppermann, Wolf Bechstein, Martin-Leo Hans­
mann, Ralf Küppers, and Bräuninger Andreas. 2009. “Inactivating SOCS1 Mutations are Caused
by Aberrant Somatic Hypermutation and Restricted to a Subset of B-Cell Lymphoma Entities.”
Blood 114 (20): 4503–6. doi: 10.1182/blood-2009-06-225839.
Mottok, Anja, Christoph Renné, Klaus Willenbrock, Martin-Leo Hansmann, and Bräuninger Andreas.
2007. “Somatic Hypermutation of SOCS1 in Lymphocyte-Predominant Hodgkin Lymphoma Is
Accompanied by High JAK2 Expression and Activation of STAT6.” Blood 110 (9): 3387–90. doi:
10.1182/blood-2007-03-082511.
Mullighan, Charles G., Richard C. Jinghui Zhang, Harvey J. Racquel Collins-Underwood, Brenda
A. Schulman, Letha A. Phillips, Sarah K. Tasian, et al. 2009a. “JAK Mutations in High-Risk
Childhood Acute Lymphoblastic Leukemia.” Proceedings of the National Academy of Sciences 106
(23): 9414–18. doi: 10.1073/pnas.0811761106.
Mullighan, Charles G., J. Racquel Collins-Underwood, A. A. Letha, Michael G. Phillips, Wei Liu Loudin,
Jinghui Zhang, Jing Ma, et al. 2009b. “Rearrangement of CRLF2 in B-Progenitor– And down
Syndrome–Associated Acute Lymphoblastic Leukemia.” Nature Genetics 41 (11): 1243–6. doi:
10.1038/ng.469.
Mullighan, Charles G., Su Xiaoping, Jinghui Zhang, Letha A. Ina Radtke, A. Phillips, Christopher
B. Miller, Ma Jing, et al. 2009c. “Deletion of IKZF1 and Prognosis in Acute Lymphoblastic
Leukemia.” The New England Journal of Medicine 360 (5): 470–80. doi: 10.1056/NEJMoa0808253.
Nairismägi, M.-L., M. E. Gerritsen, Z. M. Li, G. C. Wijaya, B. K. H. Chia, Y. Laurensia, J. Q. Lim, et al.
2018. “Oncogenic Activation of JAK3-STAT Signaling Confers Clinical Sensitivity to PRN371,
a Novel Selective and Potent JAK3 Inhibitor, in Natural Killer/T-Cell Lymphoma.” Leukemia 32
(5): 1147–56. doi: 10.1038/s41375-017-0004-x.
Nangalia, J., C. E. Massie, E. J. Baxter, F. L. Nice, G. Gundem, D. C. Wedge, E. Avezov, et al. 2013.
“Somatic CALR Mutations in Myeloproliferative Neoplasms with Nonmutated JAK2.” The New
England Journal of Medicine 369 (25): 2391–405. doi: 10.1056/NEJMoa1312542.
Neste, Eric, Van den, Marc André, Thomas Gastinne, Aspasia Stamatoullas, Corinne Haioun,
Amine Belhabri, Oumedaly Reman, et al. 2018. “Phase II Study of Oral JAK1/JAK2 Inhibitor
Ruxolitinib in Advanced Relapsed/Refractory Hodgkin Lymphoma.” Haematologica, January,
haematol.2017.180554. doi: 10.3324/haematol.2017.180554.
Ngo, Vu N., Ryan M. Young, Roland Schmitz, Sameer Jhavar, Wenming Xiao, Kian-Huat Lim,
Holger Kohlhammer, et al. 2011. “Oncogenically Active MYD88 Mutations in Human
Lymphoma.” Nature 470 (7332): 115–19. doi: 10.1038/nature09671.
Nguyen, Melody H.-H., M.-Y. Ho Jenny, Bryan K. Beattie, and Dwayne L. Barber. 2001. “TEL-JAK2
Mediates Constitutive Activation of the Phosphatidylinositol 3′-Kinase/protein Kinase B Signaling
Pathway.” Journal of Biological Chemistry 276 (35): 32704–13. doi: 10.1074/jbc.M103100200.
Nielsen, Camilla, Henrik S. Birgens, Børge G. Nordestgaard, Lasse Kjær, and Stig E. Bojesen. 2011. “The
JAK2 V617F Somatic Mutation, Mortality and Cancer Risk in the General Population.” Haema­
tologica 96 (3): 450–3. doi: 10.3324/haematol.2010.033191.
Noone, A. M., N. Howlader, M. Krapcho, D. Miller, Brest Brest, M. Yu, J. Ruhl, et al. 2018. “SEER
Cancer Statistics Review.” National Cancer Institute, Bethesda, MD May, 1975–2015.
Norton, Alice, Chris Fisher, Hui Liu, Qiang Wen, Gina Mundschau, Jose Luis Fuster, Henrik Hasle, et al.
2007. “Analysis of JAK3, JAK2, and C-MPL Mutations in Transient Myeloproliferative Disorder
and Myeloid Leukemia of down Syndrome Blasts in Children with down Syndrome.” Blood 110 (3):
1077–9. doi: 10.1182/blood-2007-03-080374.
Nussenzveig, Roberto H., Sabina I. Swierczek, Jaroslav Jelinek, Amos Gaikwad, Enli Liu, Jaroslav
F. Prchal Srdan Verstovsek, and Josef T. Prchal. 2007. “Polycythemia Vera Is Not Initiated by
JAK2V617F Mutation.” Experimental Hematology 35 (1): 32.e1–32.e9. doi: 10.1016/j.
exphem.2006.11.012.
Oh, Stephen T., Erin F. Simonds, Matthew B. Carol Jones, Yury Goltsev Hale, Kenneth D. Gibbs, Jason
D. Merker, James L. Zehnder, Garry P. Nolan, and Jason Gotlib. 2010. “Novel Mutations in the
Inhibitory Adaptor Protein LNK Drive JAK–STAT Signaling in Patients with Myeloproliferative
Neoplasms.” Blood 116 (6): 988–92. doi: 10.1182/blood-2010-02-270108.
Oschlies, Ilske, Wolfram Klapper, Martin Zimmermann, Matthias Krams, Hans-Heinrich Wacker,
Birgit Burkhardt, Lana Harder, Reiner Siebert, Alfred Reiter, and Reza Parwaresch. 2006. “Diffuse
214 JAK-STAT Signaling in Diseases

Large B-Cell Lymphoma in Pediatric Patients Belongs Predominantly to the Germinal-Center Type
B-Cell Lymphomas: A Clinicopathologic Analysis of Cases Included in the German BFM (berlin­
frankfurt-münster) Multicenter Trial.” Blood 107 (10): 4047–52. doi: 10.1182/blood-2005-10-4213.
Parampalli, Yajnanarayana Sowmya, Thomas Stübig, Isabelle Cornez, Haefaa Alchalby,
Kathrin Schönberg, Janna Rudolph, Ioanna Triviai, et al. 2015. “JAK1/2 Inhibition Impairs
T Cell Function in Vitro and in Patients with Myeloproliferative Neoplasms.” British Journal of
Haematology 169 (6): 824–33. doi: 10.1111/bjh.13373.
Pardanani, A., T. Lasho, C. Finke, S. T. Oh, J. Gotlib, and A. Tefferi. 2010. “LNK Mutation Studies in
Blast-Phase Myeloproliferative Neoplasms, and in Chronic-Phase Disease with TET2, IDH, JAK2
or MPL Mutations.” Leukemia 24 (10): 1713–18. doi: 10.1038/leu.2010.163.
Pardanani, A., T. L. Lasho, C. Finke, C. A. Hanson, and A. Tefferi. 2007. “Prevalence and Clinicopatho­
logic Correlates of JAK2 Exon 12 Mutations in JAK2V617F-Negative Polycythemia Vera.”
Leukemia 21 (9): 1960–3. doi: 10.1038/sj.leu.2404810.
Pardanani, Animesh, Jorge E. Claire Harrison, Francisco Cervantes Cortes, Ruben A. Mesa,
Donald Milligan, Tamás Masszi, et al. 2015. “Safety and Efficacy of Fedratinib in Patients with
Primary or Secondary Myelofibrosis: A Randomized Clinical Trial.” JAMA Oncology 1 (5): 643–
51. doi: 10.1001/jamaoncol.2015.1590.
Pardanani, Animesh D., Ross L. Levine, Terra Lasho, Ruben A. Yana Pikman, Martha Wadleigh Mesa,
David P. Steensma, et al. 2006. “MPL515 Mutations in Myeloproliferative and Other Myeloid
Disorders: A Study of 1182 Patients.” Blood 108 (10): 3472–6. doi: 10.1182/blood-2006-04-018879.
Passamonti, Francesco, Chiara Elena, Radek C. Susanne Schnittger, Anthony R. Skoda, François
Girodon Green, Jean-Jacques Kiladjian, et al. 2011. “Molecular and Clinical Features of the
Myeloproliferative Neoplasm Associated with JAK2 Exon 12 Mutations.” Blood 117 (10): 2813–16.
doi: 10.1182/blood-2010-11-316810.
Passamonti, Francesco, Martin Griesshammer, Francesca Palandri, Miklos Egyed, Giulia Benevolo,
Timothy Devos, Jeannie Callum, et al. 2017. “Ruxolitinib for the Treatment of Inadequately Controlled
Polycythaemia Vera without Splenomegaly (RESPONSE-2): A Randomised, Open-Label, Phase 3b
Study.” The Lancet Oncology 18 (1): 88–99. doi: 10.1016/S1470-2045(16)30558-7.
Peeters, Pieter, Sophie D. Raynaud, Jan Cools, Iwona Wlodarska, Josiane Grosgeorge, Patrick Philip,
Fabrice Monpoux, et al. 1997. “Fusion of TEL, the ETS-Variant Gene 6 (ETV6), to the
Receptor-Associated Kinase JAK2 as a Result of T(9; 12) in a Lymphoid and T(9; 15; 12) in
a Myeloid Leukemia.” Blood 90 (7): 2535–40. www.bloodjournal.org/content/90/7/2535.
Pemmaraju, Naveen, Hagop Kantarjian, Tapan Kadia, Jorge Cortes, Gautam Borthakur, Kate Newberry,
Guillermo Garcia-Manero, et al. 2015. “A Phase I/II Study of the Janus Kinase (JAK)1 and 2
Inhibitor Ruxolitinib in Patients with Relapsed or Refractory Acute Myeloid Leukemia.” Clinical
Lymphoma Myeloma and Leukemia 15 (3): 171–6. doi: 10.1016/j.clml.2014.08.003.
Perez-Andreu, Virginia, Kathryn G. Roberts, Richard C. Harvey, Wenjian Yang, Cheng Cheng,
Deqing Pei, Xu Heng, et al. 2013. “Inherited GATA3 Variants are Associated with Ph-like Child­
hood Acute Lymphoblastic Leukemia and Risk of Relapse.” Nature Genetics 45 (12): 1494–8. doi:
10.1038/ng.2803.
Perez-Andreu, Virginia, Kathryn G. Roberts, Xu Heng, Colton Smith, Hui Zhang, Wenjian Yang,
Richard C. Harvey, et al. 2015. “A Genome-Wide Association Study of Susceptibility to Acute
Lymphoblastic Leukemia in Adolescents and Young Adults.” Blood 125 (4): 680–6. doi: 10.1182/
blood-2014-09-595744.
Pietra, Daniela, Matteo C. Ilaria Casetti, Chiara Da Vià, Chiara Milanesi Elena, and Elisa Rumi. 2012.
“JAK2 GGCC Haplotype in MPL Mutated Myeloproliferative Neoplasms.” American Journal of
Hematology 87 (7): 746–7. doi: 10.1002/ajh.23229.
Pietra, Daniela, Li Sai, Angela Brisci, Francesco Passamonti, Elisa Rumi, Alexandre Theocharides,
Maurizio Ferrari, et al. 2008. “Somatic Mutations of JAK2 Exon 12 in Patients with JAK2
(v617f)-negative Myeloproliferative Disorders.” Blood 111 (3): 1686–9. doi: 10.1182/blood-2007­
07-101576.
Pikman, Yana, Benjamin H. Lee, Thomas Mercher, Benjamin L. Elizabeth McDowell, Maricel
Gozo Ebert, Adam Cuker, et al. 2006. “MPLW515L Is a Novel Somatic Activating Mutation in
Myelofibrosis with Myeloid Metaplasia.” PLoS Medicine 3 (7): e270. doi: 10.1371/journal.
pmed.0030270.
JAK-STAT Signaling in Hematologic Malignancies 215

Pillai, Vinodh, Michael Tallarico, Michael R. Bishop, and Megan S. Lim. 2016. “Mature T- and NK-Cell
Non-Hodgkin Lymphoma in Children and Young Adolescents.” British Journal of Haematology
173 (4): 573–81. doi: 10.1111/bjh.14044.
Porcu, Michaël, Olga Gielen, Jan Cools, and De Keersmaecker Kim. 2009. “JAK1 Mutation Analysis in
T-Cell Acute Lymphoblastic Leukemia Cell Lines.” Haematologica 94 (3): 435–7. doi: 10.3324/
haematol.13587.
Prutsch, Nicole, Elisabeth Gurnhofer, Tobias Suske, Huan Chang Liang, Michaela Schlederer,
Simone Roos, Lawren C. Wu, et al. 2018. “Dependency on the TYK2/STAT1/MCL1 Axis in
Anaplastic Large Cell Lymphoma.” Leukemia, August, 1. doi: 10.1038/s41375-018-0239-1.
Pui, Ching-Hon, Kathryn G. Roberts, Jun J. Yang, and Charles G. Mullighan. 2017. “Philadelphia
Chromosome-Like Acute Lymphoblastic Leukemia.” Clinical Lymphoma, Myeloma & Leukemia
17 (8): 464–70. doi: 10.1016/j.clml.2017.03.299.
Putti, Maria Caterina, Marco Pizzi, Irene Bertozzi, Elena Sabattini, Concetta Micalizzi, Piero Farruggia,
Ugo Ramenghi, et al. 2017. “Bone Marrow Histology for the Diagnosis of Essential Thrombo­
cythemia in Children: A Multicenter Italian Study.” Blood 129 (22): 3040–42. doi: 10.1182/blood­
2017-01-761767.
Qin, Haiying, Monica Cho, Waleed Haso, Sarah K. Ling Zhang, Htoo Tasian, Oo Zarni, Gian Luca Negri,
et al. 2015. “Eradication of B-ALL Using Chimeric Antigen Receptor–Expressing T Cells Targeting
the TSLPR Oncoprotein.” Blood 126 (5): 629–39. doi: 10.1182/blood-2014-11-612903.
Rampal, Raajit, Fatima Al-Shahrour, Omar Abdel-Wahab, Jay P. Patel, Jean-Philippe Brunel, Craig
H. Mermel, AdamJ. Bass, et al. 2014. “Integrated Genomic Analysis Illustrates the Central Role of
JAK–STAT Pathway Activation in Myeloproliferative Neoplasm Pathogenesis.” Blood 123 (22):
e123–33. doi: 10.1182/blood-2014-02-554634.
Randi, Maria L., Giulia Geranio, Irene Bertozzi, Concetta Micalizzi, Ugo Ramenghi, Fabio Tucci, Lucia
D. Notarangelo, et al. 2015. “Are All Cases of Paediatric Essential Thrombocythaemia Really
Myeloproliferative Neoplasms? Analysis of a Large Cohort.” British Journal of Haematology 169
(4): 584–9. doi: 10.1111/bjh.13329.
Redell, Michele S., Marcos J. Ruiz, Todd A. Alonzo, Robert B. Gerbing, and David J. Tweardy. 2011.
“Stat3 Signaling in Acute Myeloid Leukemia: Ligand-Dependent and -independent Activation and
Induction of Apoptosis by a Novel Small-Molecule Stat3 Inhibitor.” Blood 117 (21): 5701–9. doi:
10.1182/blood-2010-04-280123.
Redell, Michele S., Marcos J. Ruiz, Robert B. Gerbing, Todd A. Alonzo, Beverly J. Lange, David
J. Tweardy, and Soheil Meshinchi. 2013. “FACS Analysis of Stat3/5 Signaling Reveals Sensitivity
to G-CSF and IL-6 as A Significant Prognostic Factor in Pediatric AML: A Children’s Oncology
Group Report.” Blood 121 (7): 1083–93. doi: 10.1182/blood-2012-04-421925.
Reshmi, Shalini C., Richard C. Harvey, Kathryn G. Roberts, Eileen Stonerock, Amy Smith,
Heather Jenkins, I.-Ming Chen, et al. 2017. “Targetable Kinase Gene Fusions in High-Risk
B-ALL: A Study from the Children’s Oncology Group.” Blood 129 (25): 3352–61. doi: 10.1182/
blood-2016-12-758979.
Ritz, Olga, Chrystelle Guiter, Flavia Castellano, Karola Dorsch, Julia Melzner, Jean-Philippe Jais,
Gwendoline Dubois, Philippe Gaulard, Peter Möller, and Karen Leroy. 2009. “Recurrent Muta­
tions of the STAT6 DNA Binding Domain in Primary Mediastinal B-Cell Lymphoma.” Blood 114
(6): 1236–42. doi: 10.1182/blood-2009-03-209759.
Roberts, Kathryn G., Ryan D. Morin, Jinghui Zhang, Martin Hirst, Yongjun Zhao, Su Xiaoping, Shann-
Ching Chen, et al. 2012. “Genetic Alterations Activating Kinase and Cytokine Receptor Signaling
in High-Risk Acute Lymphoblastic Leukemia.” Cancer Cell 22 (2): 153–66. doi: 10.1016/j.
ccr.2012.06.005.
Roberts, Kathryn G., Shalini C. Reshmi, Richard C. Harvey, I.-Ming Chen, Kinnari Patel,
Eileen Stonerock, Heather Jenkins, et al. 2018. “Genomic and Outcome Analyses of Ph-like ALL
in NCI Standard-Risk Patients: A Report from the Children’s Oncology Group.” Blood 132 (8):
815–24. doi: 10.1182/blood-2018-04-841676.
Roberts, Kathryn G., Yung-Li Yang, Debbie Payne-Turner, Jacob K. Wenwei Lin, Kirsten Dickerson Files,
Zhaohui Gu, et al. 2017. “Oncogenic Role and Therapeutic Targeting of ABL-Class and JAK–STAT
Activating Kinase Alterations in Ph-like ALL.” Blood Advances 1 (20): 1657–71. doi: 10.1182/
bloodadvances.2017011296.
216 JAK-STAT Signaling in Diseases

Roberts, Kathryn G., Li Yongjin, Debbie Payne-Turner, Richard C. Harvey, Yung-Li Yang, Deqing Pei,
Kelly McCastlain, et al. 2014. “Targetable Kinase-Activating Lesions in Ph-like Acute Lympho­
blastic Leukemia.” The New England Journal of Medicine 371 (11): 1005–15. doi: 10.1056/
NEJMoa1403088.
Roberts, Kathryn G., Gu Zhaohui, Debbie Payne-Turner, Richard C. Kelly McCastlain, I-Ming
Chen Harvey, Deqing Pei, et al. 2016. “High Frequency and Poor Outcome of Philadelphia
Chromosome–Like Acute Lymphoblastic Leukemia in Adults.” Journal of Clinical Oncology 35
(4): 394–401. doi: 10.1200/JCO.2016.69.0073.
Roemer, Margaretha G.M., Ranjana H. Advani, Azra H. Ligon, Robert A. Yasodha Natkunam, Heather
Homer Redd, Courtney F. Connelly, et al. 2016. “PD-L1 and PD-L2 Genetic Alterations Define
Classical Hodgkin Lymphoma and Predict Outcome.” Journal of Clinical Oncology 34 (23): 2690–7.
doi: 10.1200/JCO.2016.66.4482.
Rooij, Jasmijn D. E. de, C. Michel Zwaan, and Marry van den Heuvel-Eibrink. 2015. “Pediatric AML:
From Biology to Clinical Management.” Journal of Clinical Medicine 4 (1): 127–49. doi: 10.3390/
jcm4010127.
Roosbroeck, Katrien Van, Luk Cox, Thomas Tousseyn, Idoya Lahortiga, Olga Gielen,
Barbara Cauwelier, Pascale De Paepe, et al. 2011. “JAK2 Rearrangements, Including the Novel
SEC31A-JAK2 Fusion, are Recurrent in Classical Hodgkin Lymphoma.” Blood 117 (15): 4056–64.
doi: 10.1182/blood-2010-06-291310.
Rosenwald, Andreas, George Wright, Karen Leroy, Yu Xin, Randy D. Philippe Gaulard, Wing
C. Chan Gascoyne, et al. 2003. “Molecular Diagnosis of Primary Mediastinal B Cell Lymphoma
Identifies a Clinically Favorable Subgroup of Diffuse Large B Cell Lymphoma Related to Hodgkin
Lymphoma.” Journal of Experimental Medicine 198 (6): 851–62. doi: 10.1084/jem.20031074.
Rui, Lixin, N. C. Tolga Emre, Michael J. Kruhlak, Hye-Jung Chung, Christian Steidl, Graham Slack,
George W. Wright, et al. 2010. “Cooperative Epigenetic Modulation by Cancer Amplicon Genes.”
Cancer Cell 18 (6): 590–605. doi: 10.1016/j.ccr.2010.11.013.
Rumi, Elisa, Daniela Pietra, Virginia Ferretti, Thorsten Klampfl, Ashot S. Harutyunyan, Jelena
D. Milosevic, Nicole C. C. Them, et al. 2014a. “JAK2 or CALR Mutation Status Defines Subtypes
of Essential Thrombocythemia with Substantially Different Clinical Course and Outcomes.” Blood
123 (10): 1544–51. doi: 10.1182/blood-2013-11-539098.
Rumi, Elisa, Daniela Pietra, Cristiana Pascutto, Paola Guglielmelli, Alejandra Martínez-Trillos,
Ilaria Casetti, Dolors Colomer, et al. 2014b. “Clinical Effect of Driver Mutations of JAK2,
CALR, or MPL in Primary Myelofibrosis.” Blood 124 (7): 1062–9. doi: 10.1182/blood-2014-05­
578435.
Russell, Lisa J., Melania Capasso, Inga Vater, Olivier A. Takashi Akasaka, Maria Bernard,
Jose Calasanz, Thiruppavaii Chandrasekaran, et al. 2009. “Deregulated Expression of Cytokine
Receptor Gene, CRLF2, Is Involved in Lymphoid Transformation in B-Cell Precursor Acute
Lymphoblastic Leukemia.” Blood 114 (13): 2688–98. doi: 10.1182/blood-2009-03-208397.
Saksena, Annapurna, Prerna Arora, G. R. Sethi Nita Khurana, and Tejinder Singh. 2014. “Paediatric
Idiopathic Myelofibrosis.” Indian Journal of Hematology & Blood Transfusion: an Official Journal of
Indian Society of Hematology and Blood Transfusion 30 (Suppl 1): 363–5. doi: 10.1007/s12288-014­
0412-2.
Sangkhae, Veena S., Leah Etheridge, Kenneth Kaushansky, and Ian S. Hitchcock. 2014. “The Thrombo­
poietin Receptor, MPL, Is Critical for Development of a JAK2V617F-Induced Myeloproliferative
Neoplasm.” Blood 124 (26): 3956–63. doi: 10.1182/blood-2014-07-587238.
Santos, Nuno R. dos, David S. Rickman, Aurélien de Reynies, Françoise Cormier, Maryvonne Williame,
Camille Blanchard, Marc-Henri Stern, and Jacques Ghysdael. 2007. “Pre-TCR Expression Coop­
erates with TEL-JAK2 to Transform Immature Thymocytes and Induce T-Cell Leukemia.” Blood
109 (9): 3972–81. doi: 10.1182/blood-2006-09-048801.
Sato, Tomohiko, Tsutomu Toki, Rika Kanezaki, Xu Gang, Kiminori Terui, Hirokazu Kanegane,
Masayoshi Miura, et al. 2008. “Functional Analysis of JAK3 Mutations in Transient Myeloproli­
ferative Disorder and Acute Megakaryoblastic Leukaemia Accompanying down Syndrome.”
British Journal of Haematology 141 (5): 681–8. doi: 10.1111/j.1365-2141.2008.07081.x.
Savage, Kerry J., Jeffery L. Stefano Monti, Giorgio Cattoretti Kutok, Donna Neuberg, Laurence de
Leval, Paul Kurtin, et al. 2003. “The Molecular Signature of Mediastinal Large B-Cell Lymphoma
JAK-STAT Signaling in Hematologic Malignancies 217

Differs from that of Other Diffuse Large B-Cell Lymphomas and Shares Features with Classical
Hodgkin Lymphoma.” Blood 102 (12): 3871–9. doi: 10.1182/blood-2003-06-1841.
Savino, A. M., J. Sarno, L. Trentin, M. Vieri, G. Fazio, M. Bardini, C. Bugarin, et al. 2017. “The Histone
Deacetylase Inhibitor Givinostat (ITF2357) Exhibits Potent Anti-Tumor Activity against
CRLF2-Rearranged BCP-ALL.” Leukemia 31 (11): 2365–75. doi: 10.1038/leu.2017.93.
Scheeren, Ferenc A., Anja U. van Lent, Maho Nagasawa, Kees Weijer, Hergen Spits, Nicolas Legrand,
and Bianca Blom. 2010. “Thymic Stromal Lymphopoietin Induces Early Human B-Cell Prolifera­
tion and Differentiation.” European Journal of Immunology 40 (4): 955–65. doi: 10.1002/
eji.200939419.
Schinnerl, Dagmar, Klaus Fortschegger, João R. Maximilian Kauer, M. Marchante, Monique
L. Reinhard Kofler, Den Boer, and Sabine Strehl. 2015. “The Role of the Janus-Faced Transcription
Factor PAX5-JAK2 in Acute Lymphoblastic Leukemia.” Blood 125 (8): 1282–91. doi: 10.1182/
blood-2014-04-570960.
Schmitz, Roland, George W. Wright, Calvin A. Da Wei Huang, James D. Johnson, James Q. Phelan,
Sandrine Roulland Wang, et al. 2018. “Genetics and Pathogenesis of Diffuse Large B-Cell
Lymphoma.” The New England Journal of Medicine 378 (15): 1396–407. doi: 10.1056/
NEJMoa1801445.
Schultz, K. R., A. Carroll, N. A. Heerema, W. P. Bowman, A. Aledo, W. B. Slayton, H. Sather, et al. 2014.
“Long-Term Follow-up of Imatinib in Pediatric Philadelphia Chromosome-Positive Acute Lym­
phoblastic Leukemia: Children’s Oncology Group Study AALL0031.” Leukemia 28 (7): 1467–71.
doi: 10.1038/leu.2014.30.
Schwaller, J., J. Frantsve, J. Aster, I. R. Williams, M. H. Tomasson, T. S. Ross, P. Peeters, et al. 1998.
“Transformation of Hematopoietic Cell Lines to Growth-Factor Independence and Induction of
a Fatal Myelo- and Lymphoproliferative Disease in Mice by Retrovirally Transduced TEL/JAK2
Fusion Genes.” The EMBO Journal 17 (18): 5321–33. doi: 10.1093/emboj/17.18.5321.
Schwartzman, Omer, Angela Maria Savino, Michael Gombert, Chiara Palmi, Gunnar Cario,
Martin Schrappe, Cornelia Eckert, et al. 2017. “Suppressors and Activators of JAK–STAT
Signaling at Diagnosis and Relapse of Acute Lymphoblastic Leukemia in down Syndrome.”
Proceedings of the National Academy of Sciences of the United States of America 114 (20): E4030–39.
doi: 10.1073/pnas.1702489114.
Scott, Linda M., Peter J. Campbell, E. Joanna Baxter, Tony Todd, Philip Stephens, Sarah Edkins,
Michael R. Richard Wooster, Stratton P. Andrew Futreal, and Anthony R. Green. 2005. “The
V617F JAK2 Mutation Is Uncommon in Cancers and in Myeloid Malignancies Other than the
Classic Myeloproliferative Disorders.” Blood 106 (8): 2920–1. doi: 10.1182/blood-2005-05-2087.
Scott, Linda M., Mike A. Scott, Peter J. Campbell, and Anthony R. Green. 2006. “Progenitors
Homozygous for the V617F Mutation Occur in Most Patients with Polycythemia Vera, but Not
Essential Thrombocythemia.” Blood 108 (7): 2435–7. doi: 10.1182/blood-2006-04-018259.
Scott, Linda M., Ross L. Wei Tong, Mike A. Levine, Philip A. Scott, Michael R. Beer, Stratton P. Andrew
Futreal, et al. 2007. “JAK2 Exon 12 Mutations in Polycythemia Vera and Idiopathic Erythrocytosis.”
New England Journal of Medicine 356 (5): 459–68. doi: 10.1056/NEJMoa065202.
Sekiya, Yuko, Yusuke Okuno, Hideki Muramatsu, Olfat Ismael, Nozomu Kawashima, Atsushi Narita,
Xinan Wang, et al. 2016. “JAK2, MPL and CALR Mutations in Children with Essential
Thrombocythemia.” International Journal of Hematology 104 (2): 266–7. doi: 10.1007/s12185-016­
2022-2.
Semenza, Gregg L. 2009. “Involvement of Oxygen-Sensing Pathways in Physiologic and Pathologic
Erythropoiesis.” Blood 114 (10): 2015–19. doi: 10.1182/blood-2009-05-189985.
Senkevitch, Emilee, Li Wenqing, Julie A. Hixon, Sarah D. Caroline Andrews, Gary T. Cramer, Timothy
Back Pauly, Kelli Czarra, and Scott K. Durum. 2018. “Inhibiting Janus Kinase 1 and BCL-2 to
Treat T Cell Acute Lymphoblastic Leukemia with IL7-Rα Mutations.” Oncotarget 9 (32): 22605–17.
doi: 10.18632/oncotarget.25194.
Shah, Niketa, Ali Al-Ahmari, Arwa Al-Yamani, Lee Dupuis, Derek Stephens, and Johann Hitzler. 2009.
“Outcome and Toxicity of Chemotherapy for Acute Lymphoblastic Leukemia in Children with
Down Syndrome.” Pediatric Blood & Cancer 52 (1): 14–19. doi: 10.1002/pbc.21737.
Shochat, Chen, Obul R. Noa Tal, Chiara Palmi Bandapalli, Ithamar Ganmore, Geertruy Te Kronnie,
Gunnar Cario, et al. 2011. “Gain-of-Function Mutations in Interleukin-7 Receptor-α (IL7R) in
218 JAK-STAT Signaling in Diseases

Childhood Acute Lymphoblastic Leukemias.” Journal of Experimental Medicine 208 (5): 901–8.
doi: 10.1084/jem.20110580.
Sidon, P., H. El Housni, B. Dessars, and P. Heimann. 2006. “The JAK2V617F Mutation Is Detectable at
Very Low Level in Peripheral Blood of Healthy Donors.” Leukemia 20 (9): 1622. doi: 10.1038/sj.
leu.2404292.
Silva, M., C. Richard, A. Benito, C. Sanz, I. Olalla, and J. L. Fernández-Luna. 1998. “Expression of
Bcl-x in Erythroid Precursors from Patients with Polycythemia Vera.” The New England Journal of
Medicine 338 (9): 564–71. doi: 10.1056/NEJM199802263380902.
Silvennoinen, Olli, and Stevan R. Hubbard. 2015. “Molecular Insights into Regulation of JAK2 in
Myeloproliferative Neoplasms.” Blood 125 (22): 3388–92. doi: 10.1182/blood-2015-01-621110.
Skoda, Radek C. 2010. “Hereditary Myeloproliferative Disorders.” Haematologica 95 (1): 6–8. doi:
10.3324/haematol.2009.015941.
Slayton, William B., Kirk R. Schultz, John A. Kairalla, Meenakshi Devidas, Mi Xinlei, Michael
A. Pulsipher, Bill H. Chang, et al. 2018. “Dasatinib Plus Intensive Chemotherapy in Children,
Adolescents, and Young Adults with Philadelphia Chromosome-Positive Acute Lymphoblastic
Leukemia: Results of Children’s Oncology Group Trial AALL0622.” Journal of Clinical Oncology:
Official Journal of the American Society of Clinical Oncology 36 (22): 2306–14. doi: 10.1200/
JCO.2017.76.7228.
Smith, Malcolm A., Nita L. Seibel, Sean F. Altekruse, A. G. Lynn, Danielle L. Ries, Maura
O’Leary Melbert, Franklin O. Smith, and Gregory H. Reaman. 2010. “Outcomes for Children
and Adolescents with Cancer: Challenges for the Twenty-First Century.” Journal of Clinical
Oncology 28 (15): 2625–34. doi: 10.1200/JCO.2009.27.0421.
Song, Tammy Linlin, Maarja-Liisa Nairismägi, Yurike Laurensia, Jing-Quan Lim, Jing Tan, Zhi-Mei Li,
Wan-Lu Pang, et al. 2018. “Oncogenic Activation of STAT3 Pathway Drives PD-L1 Expression in
Natural Killer/T Cell Lymphoma.” Blood, January, blood-2018-01-829424. doi: 10.1182/blood­
2018-01-829424.
Springuel, Lorraine, Tekla Hornakova, Elisabeth Losdyck, Fanny Lambert, Stefan N. Emilie Leroy,
Elisabetta Flex Constantinescu, Marco Tartaglia, Laurent Knoops, and Jean-Christophe Renauld.
2014. “Cooperating JAK1 and JAK3 Mutants Increase Resistance to JAK Inhibitors.” Blood 124
(26): 3924–31. doi: 10.1182/blood-2014-05-576652.
Staerk, Judith, Anders Kallin, Jean-Baptiste Demoulin, William Vainchenker, and Stefan
N. Constantinescu. 2005. “JAK1 and Tyk2 Activation by the Homologous Polycythemia Vera
JAK2 V617F Mutation CROSS-TALK WITH IGF1 RECEPTOR.” Journal of Biological Chem­
istry 280 (51): 41893–9. doi: 10.1074/jbc.C500358200.
Staerk, Judith, Catherine Lacout, Steven O. Takeshi Sato, William Vainchenker Smith, and Stefan
N. Constantinescu. 2006. “An Amphipathic Motif at the Transmembrane-Cytoplasmic Junction
Prevents Autonomous Activation of the Thrombopoietin Receptor.” Blood 107 (5): 1864–71. doi:
10.1182/blood-2005-06-2600.
Steensma, David P., Gordon W. Dewald, Terra L. Lasho, Heather L. Powell, Rebecca F. McClure, Ross
L. Levine, D. Gary Gilliland, and Ayalew Tefferi. 2005. “The JAK2 V617F Activating Tyrosine
Kinase Mutation Is an Infrequent Event in Both ‘Atypical’ Myeloproliferative Disorders and
Myelodysplastic Syndromes.” Blood 106 (4): 1207–9. doi: 10.1182/blood-2005-03-1183.
Steensma, D. P., R. F. McClure, J. E. Karp, A. Tefferi, T. L. Lasho, H. L. Powell, G. W. DeWald, and
S. H. Kaufmann. 2006. “JAK2 V617F Is a Rare Finding in De Novo Acute Myeloid Leukemia, but
STAT3 Activation Is Common and Remains Unexplained.” Leukemia 20 (6): 971–8. doi: 10.1038/
sj.leu.2404206.
Stevens, Alexandra M., Marcos J. Ruiz, Robert B. Gerbing, Todd A. Alonzo, Alan S. Gamis, and Michele
S. Redell. 2015. “Ligand-Induced STAT3 Signaling Increases at Relapse and Is Associated with
Outcome in Pediatric Acute Myeloid Leukemia: A Report from the Children’s Oncology Group.”
Haematologica 100 (12): e496–500. doi: 10.3324/haematol.2015.131508.
Suryani, Santi, S. Lauryn, Richard C. Bracken, Keith C. Harvey, S. Sia Hernan Carol, I.-Ming Chen,
Kathryn Evans, et al. 2015. “Evaluation of the In Vitro and In Vivo Efficacy of the JAK Inhibitor
AZD1480 against JAK–Mutated Acute Lymphoblastic Leukemia.” Molecular Cancer Therapeutics
14 (2): 364–74. doi: 10.1158/1535-7163.MCT-14-0647.
JAK-STAT Signaling in Hematologic Malignancies 219

Swerdlow, S. H., E. Campo, N. L. Harris, E. S. Jaffe, S. A. Pileri, H. Stein, J. Thiele, and J. W. Vardiman.
2008. “WHO Classification of Tumours of Haematopoietic and Lymphoid Tissues.” IARC Pub­
lications. http://publications.iarc.fr/Book-And-Report-Series/Who-Iarc-Classification-Of-Tumours/
WHO-Classification-Of-Tumours-Of-Haematopoietic-And-Lymphoid-Tissues-2008.
Swerdlow, Steven H., Stefano A. Elias Campo, Nancy Pileri, Lee Harris, Harald Stein, Reiner Siebert,
Ranjana Advani, et al. 2016. “The 2016 Revision of the World Health Organization Classification
of Lymphoid Neoplasms.” Blood 127 (20): 2375–90. doi: 10.1182/blood-2016-01-643569.
Szpurka, H., L. P. Gondek, S. R. Mohan, E. D. Hsi, K. S. Theil, and J. P. Maciejewski. 2009. “UPD1p
Indicates the Presence of MPL W515L Mutation in RARS-T, a Mechanism Analogous to UPD9p
and JAK2 V617F Mutation.” Leukemia 23 (3): 610–14. doi: 10.1038/leu.2008.249.
Tasian, Sarah K., Deborah S. Albert Assad, Yining Du Hunter, and Mignon L. Loh. 2018. “A Phase 2
Study of Ruxolitinib with Chemotherapy in Children with Philadelphia Chromosome-Like Acute
Lymphoblastic Leukemia (INCB18424-269/AALL1521): Dose-Finding Results from the Part 1
Safety Phase.” Blood 132 (Suppl 1): 555. doi: 10.1182/blood-2018-99-110221.
Tasian, Sarah K., Michelle Y. Doral, Michael J. Borowitz, Brent L. Wood, I.-Ming Chen, Richard
C. Harvey, Julie M. Gastier-Foster, et al. 2012. “Aberrant STAT5 and PI3K/MTOR Pathway
Signaling Occurs in Human CRLF2-Rearranged B-Precursor Acute Lymphoblastic Leukemia.”
Blood 120 (4): 833–42. doi: 10.1182/blood-2011-12-389932.
Tasian, Sarah K., David T. Teachey, Li Yong, Feng Shen, Richard C. Harvey, I.-Ming Chen,
Theresa Ryan, et al. 2017. “Potent Efficacy of Combined PI3K/MTOR and JAK or ABL Inhibition
in Murine Xenograft Models of Ph-like Acute Lymphoblastic Leukemia.” Blood 129 (2): 177–87.
doi: 10.1182/blood-2016-05-707653.
Tefferi, A. 2010. “Novel Mutations and Their Functional and Clinical Relevance in Myeloproliferative
Neoplasms: JAK2, MPL, TET2, ASXL1, CBL, IDH and IKZF1.” Leukemia 24 (6): 1128–38. doi:
10.1038/leu.2010.69.
Tefferi, A., T. L. Lasho, C. M. Finke, R. A. Knudson, R. Ketterling, C. H. Hanson, M. Maffioli,
D. Caramazza, F. Passamonti, and A. Pardanani. 2014a. “CALR Vs JAK2 Vs MPL-Mutated or
Triple-Negative Myelofibrosis: Clinical, Cytogenetic and Molecular Comparisons.” Leukemia 28
(7): 1472–7. doi: 10.1038/leu.2014.3.
Tefferi, A., E. Rumi, G. Finazzi, H. Gisslinger, A. M. Vannucchi, F. Rodeghiero, M. L. Randi, et al. 2013.
“Survival and Prognosis among 1545 Patients with Contemporary Polycythemia Vera: An Interna­
tional Study.” Leukemia 27 (9): 1874–81. doi: 10.1038/leu.2013.163.
Tefferi, Ayalew. 2006. “Classification, Diagnosis and Management of Myeloproliferative Disorders in the
JAK2V617F Era.” ASH Education Program Book 2006 (1): 240–5. doi: 10.1182/asheducation­
2006.1.240.
Tefferi, Ayalew, Juergen Thiele, Attilio Orazi, Hans Michael Kvasnicka, Curtis A. Tiziano Barbui,
Giovanni Barosi Hanson, et al. 2007. “Proposals and Rationale for Revision of the World Health
Organization Diagnostic Criteria for Polycythemia Vera, Essential Thrombocythemia, and Primary
Myelofibrosis: Recommendations from an Ad Hoc International Expert Panel.” Blood 110 (4):
1092–97. doi: 10.1182/blood-2007-04-083501.
Tefferi, Ayalew, Rakhee Vaidya, Domenica Caramazza, Christy Finke, Terra Lasho, and
Animesh Pardanani. 2011. “Circulating Interleukin (IL)-8, IL-2R, IL-12, and IL-15 Levels Are
Independently Prognostic in Primary Myelofibrosis: A Comprehensive Cytokine Profiling Study.”
Journal of Clinical Oncology 29 (10): 1356–63. doi: 10.1200/JCO.2010.32.9490.
Tefferi, Ayalew, and William Vainchenker. 2011. “Myeloproliferative Neoplasms: Molecular Pathophy­
siology, Essential Clinical Understanding, and Treatment Strategies.” Journal of Clinical Oncology
29 (5): 573–82. doi: 10.1200/JCO.2010.29.8711.
Tefferi, Ayalew, Emnet A. Wassie, Paola Guglielmelli, Alem A. Naseema Gangat, Terra L. Belachew,
Christy Finke Lasho, et al. 2014b. “Type 1 versus Type 2 Calreticulin Mutations in Essential
Thrombocythemia: A Collaborative Study of 1027 Patients.” American Journal of Hematology 89
(8): E121–4. doi: 10.1002/ajh.23743.
Teofili, Luciana, Tonia Cenci, Maurizio Martini, Sara Capodimonti, Lorenza Torti, Fiorina Giona,
Angela Amendola, et al. 2009. “The Mutant JAK2V617F Allele Burden in Children with Essential
Thrombocythemia.” British Journal of Haematology 145 (3): 430–2. doi: 10.1111/j.1365­
2141.2009.07591.x.
220 JAK-STAT Signaling in Diseases

Teofili, Luciana, Robin Foà, Fiorina Giona, and Luigi Maria Larocca. 2008. “Childhood Polycythemia
Vera and Essential Thrombocythemia: Does Their Pathogenesis Overlap with that of Adult
Patients?” Haematologica 93 (2): 169–72. doi: 10.3324/haematol.12002.
Teofili, Luciana, Fiorina Giona, Maurizio Martini, Tonia Cenci, Francesco Guidi, Lorenza Torti,
Giovanna Palumbo, Angela Amendola, Robin Foà, and Luigi M. Larocca. 2007a. “Markers of
Myeloproliferative Diseases in Childhood Polycythemia Vera and Essential Thrombocythemia.”
Journal of Clinical Oncology 25 (9): 1048–53. doi: 10.1200/JCO.2006.08.6884.
Teofili, Luciana, Fiorina Giona, Maurizio Martini, Tonia Cenci, Francesco Guidi, Lorenza Torti,
Giovanna Palumbo, Angela Amendola, Giuseppe Leone, et al. 2007b. “The Revised WHO
Diagnostic Criteria for Ph-Negative Myeloproliferative Diseases are Not Appropriate for the
Diagnostic Screening of Childhood Polycythemia Vera and Essential Thrombocythemia.” Blood
110 (9): 3384–6. doi: 10.1182/blood-2007-06-094276.
Teofili, Luciana, Fiorina Giona, Lorenza Torti, Tonia Cenci, Bianca Maria Ricerca, Carlo Rumi,
Vittorio Nunes, et al. 2010. “Hereditary Thrombocytosis Caused by MPLSer505Asn Is Associated
with a High Thrombotic Risk, Splenomegaly and Progression to Bone Marrow Fibrosis.” Haema­
tologica 95 (1): 65–70. doi: 10.3324/haematol.2009.007542.
Teofili, Luciana, Maurizio Martini, Tonia Cenci, Giovanna Petrucci, Lorenza Torti, Sergio Storti,
Francesco Guidi, Giuseppe Leone, and Luigi Maria Larocca. 2007c. “Different STAT-3 and
STAT-5 Phosphorylation Discriminates among Ph-Negative Chronic Myeloproliferative Diseases
and Is Independent of the V617F JAK–2 Mutation.” Blood 110 (1): 354–9. doi: 10.1182/blood­
2007-01-069237.
Thomson, A. B., and W. H. B. Wallace. 2002. “Treatment of Paediatric Hodgkin’s Disease. A Balance of
Risks.” European Journal of Cancer (Oxford, England: 1990) 38 (4): 468–77.
Tiacci, Enrico, Erik Ladewig, Gianluca Schiavoni, Alex Penson, Elisabetta Fortini, Valentina Pettirossi,
Yuchun Wang, et al. 2018. “Pervasive Mutations of JAK–STAT Pathway Genes in Classical
Hodgkin Lymphoma.” Blood 131 (22): 2454–65. doi: 10.1182/blood-2017-11-814913.
Tiedt, Ralph, Marta A. Hui Hao-Shen, Renate Looser Sobas, Stephan Dirnhofer, Jürg Schwaller, and
Radek C. Skoda. 2008. “Ratio of Mutant JAK2-V617F to Wild-Type Jak2 Determines the MPD
Phenotypes in Transgenic Mice.” Blood 111 (8): 3931–40. doi: 10.1182/blood-2007-08-107748.
Tirado, Carlos A., Weina Chen, Lily Jun-shen Huang, Matthew C. Carrie Laborde, Federico
Valdez Hiemenz, Ho Kevin, Naomi Winick, Zhenjun Lou, and Prasad Koduru. 2010. “Novel
JAK2 Rearrangement Resulting from a t(9;22)(P24;Q11.2) In B-Acute Lymphoblastic Leukemia.”
Leukemia Research 34 (12): 1674–6. doi: 10.1016/j.leukres.2010.05.031.
Tomasson, Michael H., Zhifu Xiang, Richard Walgren, Yu Zhao, Yumi Kasai, Tracie Miner, Rhonda
E. Ries, et al. 2008. “Somatic Mutations and Germline Sequence Variants in the Expressed Tyrosine
Kinase Genes of Patients with De Novo Acute Myeloid Leukemia.” Blood 111 (9): 4797–808. doi:
10.1182/blood-2007-09-113027.
Tortolani, P. J., J. A. Johnston, C. M. Bacon, D. W. McVicar, A. Shimosaka, D. Linnekin, D. L. Longo,
and J. J. O’Shea. 1995. “Thrombopoietin Induces Tyrosine Phosphorylation and Activation of the
Janus Kinase, JAK2.” Blood 85 (12): 3444–51. www.bloodjournal.org/content/85/12/3444.
Tran, Thai Hoa, Marian H. Harris, Jonathan V. Nguyen, Traci M. Blonquist, Kristen E. Stevenson,
Eileen Stonerock, Barbara L. Asselin, et al. 2018. “Prognostic Impact of Kinase-Activating Fusions
and IKZF1 Deletions in Pediatric High-Risk B-Lineage Acute Lymphoblastic Leukemia.” Blood
Advances 2 (5): 529–33. doi: 10.1182/bloodadvances.2017014704.
Treon, Steven P., Xu Lian, Guang Yang, Yangsheng Zhou, Xia Liu, Yang Cao, Patricia Sheehy, et al.
2012. “MYD88 L265P Somatic Mutation in Waldenström’s Macroglobulinemia.” New England
Journal of Medicine 367 (9): 826–33. doi: 10.1056/NEJMoa1200710.
Ugo, Valérie, Christophe Marzac, Irène Teyssandier, Frédéric Larbret, Yann Lécluse, Najet Debili,
William Vainchenker, and Nicole Casadevall. 2004. “Multiple Signaling Pathways Are Involved in
Erythropoietin-Independent Differentiation of Erythroid Progenitors in Polycythemia Vera.”
Experimental Hematology 32 (2): 179–87. doi: 10.1016/j.exphem.2003.11.003.
Ungureanu, Daniela, Wu Jinhua, Tuija Pekkala, Yashavanthi Niranjan, Ole N. Clifford Young, Chong-
Feng Xu Jensen, et al. 2011. “The Pseudokinase Domain of JAK2 Is a Dual-Specificity Protein
Kinase that Negatively Regulates Cytokine Signaling.” Nature Structural & Molecular Biology 18
(9): 971–6. doi: 10.1038/nsmb.2099.
JAK-STAT Signaling in Hematologic Malignancies 221

Ungureanu, Daniela, Pipsa Saharinen, Ilkka Junttila, Douglas J. Hilton, and Olli Silvennoinen. 2002.
“Regulation of Jak2 through the Ubiquitin-Proteasome Pathway Involves Phosphorylation of Jak2
on Y1007 and Interaction with SOCS-1.” Molecular and Cellular Biology 22 (10): 3316–26.
Vainchenker, William, Emilie Leroy, Laure Gilles, Caroline Marty, Isabelle Plo, and Stefan
N. Constantinescu. 2018. “JAK Inhibitors for the Treatment of Myeloproliferative Neoplasms
and Other Disorders.” F1000Research 7 (January): 82. doi: 10.12688/f1000research.13167.1.
Vannucchi, Alessandro M., Elisabetta Antonioli, Paola Guglielmelli, Alessandro Pancrazzi,
Vittoria Guerini, Giovanni Barosi, Marco Ruggeri, et al. 2008. “Characteristics and Clinical
Correlates of MPL 515W>L/K Mutation in Essential Thrombocythemia.” Blood 112 (3): 844–7.
doi: 10.1182/blood-2008-01-135897.
Vannucchi, Alessandro M., Jean Jacques Kiladjian, Martin Griesshammer, Tamas Masszi,
Simon Durrant, Francesco Passamonti, Claire N. Harrison, et al. 2015. “Ruxolitinib versus
Standard Therapy for the Treatment of Polycythemia Vera.” The New England Journal of Medicine
372 (5): 426–35. doi: 10.1056/NEJMoa1409002.
Vardiman, James W., Nancy Lee Harris, and Richard D. Brunning. 2002. “The World Health Organiza­
tion (WHO) Classification of the Myeloid Neoplasms.” Blood 100 (7): 2292–302. doi: 10.1182/
blood-2002-04-1199.
Velusamy, Thirunavukkarasu, Mark J. Kiel, Anagh A. Sahasrabuddhe, Delphine Rolland, Catherine
A. Dixon, Nathanael G. Bailey, Bryan L. Betz, et al. 2014. “A Novel Recurrent NPM1-TYK2 Gene
Fusion in Cutaneous CD30-Positive Lymphoproliferative Disorders.” Blood 124 (25): 3768–71. doi:
10.1182/blood-2014-07-588434.
Verstovsek, Srdan, Ruben A. Hagop Kantarjian, Animesh D. Mesa, Jorge Pardanani, Deborah
A. Cortes-Franco, Zeev Estrov Thomas, et al. 2010. “Safety and Efficacy of INCB018424, a JAK1
and JAK2 Inhibitor, in Myelofibrosis.” The New England Journal of Medicine 363 (12): 1117–27.
doi: 10.1056/NEJMoa1002028.
Verstovsek, Srdan, Francesco Passamonti, Alessandro Rambaldi, Giovanni Barosi, Elisa Rumi,
Elisabetta Gattoni, Lisa Pieri, et al. 2017. “Ruxolitinib for Essential Thrombocythemia Refractory
to or Intolerant of Hydroxyurea: Long-Term Phase 2 Study Results.” Blood 130 (15): 1768–71. doi:
10.1182/blood-2017-02-765032.
Vesely, C., C. Frech, C. Eckert, G. Cario, A. Mecklenbräuker, U. Zur Stadt, K. Nebral, et al. 2017.
“Genomic and Transcriptional Landscape of P2RY8-CRLF2-Positive Childhood Acute Lympho­
blastic Leukemia.” Leukemia 31 (7): 1491–501. doi: 10.1038/leu.2016.365.
Vita, Serena De, Claire Mulligan, Suzanne McElwaine, Franca Dagna-Bricarelli, Monica Spinelli,
Giuseppe Basso, Dean Nizetic, and Jürgen Groet. 2007. “Loss-of-Function JAK3 Mutations in
TMD and AMKL of down Syndrome.” British Journal of Haematology 137 (4): 337–41. doi:
10.1111/j.1365-2141.2007.06574.x.
Walz, Christoph, Wesam Ahmed, Katherine Lazarides, Monica Betancur, Nihal Patel,
Lothar Hennighausen, Virginia M. Zaleskas, and Richard A. Van Etten. 2012. “Essential Role for
Stat5a/b in Myeloproliferative Neoplasms Induced by BCR-ABL1 and JAK2V617F in Mice.”
Blood 119 (15): 3550–60. doi: 10.1182/blood-2011-12-397554.
Weigert, Oliver, Andrew A. Lane, Liat Bird, Nadja Kopp, Bjoern Chapuy, Diederik van Bodegom,
Angela V. Toms, et al. 2012. “Genetic Resistance to JAK2 Enzymatic Inhibitors Is Overcome by
HSP90 Inhibition.” Journal of Experimental Medicine 209 (2): 259–73. doi: 10.1084/
jem.20111694.
Weniger, M. A., I. Melzner, C. K. Menz, S. Wegener, A. J. Bucur, K. Dorsch, T. Mattfeldt, T. F. E. Barth,
and P. Möller. 2006. “Mutations of the Tumor Suppressor Gene SOCS-1 in Classical Hodgkin
Lymphoma Are Frequent and Associated with Nuclear Phospho-STAT5 Accumulation.” Oncogene
25 (18): 2679–84. doi: 10.1038/sj.onc.1209151.
Wernig, Gerlinde, Thomas Mercher, Ross L. Rachel Okabe, Benjamin H. Lee Levine, and
D. Gary Gilliland. 2006. “Expression of Jak2V617F Causes a Polycythemia Vera–Like Disease
with Associated Myelofibrosis in a Murine Bone Marrow Transplant Model.” Blood 107 (11):
4274–81. doi: 10.1182/blood-2005-12-4824.
Whitlock, James A., Harland N. Sather, Leslie L. Paul Gaynon, Robert J. Robison, Michael Trigg Wells,
Nyla A. Heerema, and Smita Bhatia. 2005. “Clinical Characteristics and Outcome of Children with
222 JAK-STAT Signaling in Diseases

Down Syndrome and Acute Lymphoblastic Leukemia: A Children’s Cancer Group Study.” Blood
106 (13): 4043–9. doi: 10.1182/blood-2003-10-3446.
Winter, Stuart S., Kimberly P. Dunsmore, Brent L. Meenakshi Devidas, Natia Esiashvili Wood,
Zhiguo Chen, Nancy Eisenberg, et al. 2018. “Improved Survival for Children and Young Adults
with T-Lineage Acute Lymphoblastic Leukemia: Results from the Children’s Oncology Group
AALL0434 Methotrexate Randomization.” Journal of Clinical Oncology: Official Journal of the
American Society of Clinical Oncology 36 (29): 2926–34. doi: 10.1200/JCO.2018.77.7250.
Wu, David, Bethany Dutra, Neal Lindeman, Hidenobu Takahashi, Nancy L. Kunihiko Takeyama,
Geraldine S. Harris, Janina Longtine Pinkus, Margaret Shipp, and Jeffery L. Kutok. 2009. “No
Evidence for the JAK2 (V617F) or JAK2 Exon 12 Mutations in Primary Mediastinal Large B-Cell
Lymphoma.” Diagnostic Molecular Pathology 18 (3): 144–9. doi: 10.1097/PDM.0b013e3181855c7f.
Wu, Shuo-Chieh, Loretta S. Li, Nadja Kopp, Joan Montero, Bjoern Chapuy, Akinori Yoda, Amanda
L. Christie, et al. 2015. “Activity of the Type II JAK2 Inhibitor CHZ868 in B Cell Acute
Lymphoblastic Leukemia.” Cancer Cell 28 (1): 29–41. doi: 10.1016/j.ccell.2015.06.005.
Xia, Z., M.R. Baer, A.W. Block, H. Baumann, and M. Wetzler. 1998. “Expression of Signal Transducers
and Activators of Transcription Proteins in Acute Myeloid Leukemia Blasts.” Cancer Research 58
(14): 3173–80.
Xiang, Zhifu, Yu Zhao, Daved H. Vesselin Mitaksov, Yumi Kasai Fremont, AnnaLynn Molitoris,
Rhonda E. Ries, et al. 2008. “Identification of Somatic JAK1 Mutations in Patients with Acute
Myeloid Leukemia.” Blood 111 (9): 4809–12. doi: 10.1182/blood-2007-05-090308.
Xu, Xuesong, Qi Zhang, Jian Luo, Shu Xing, Li Qingshan, Sanford B. Krantz, Fu Xueqi, and Zhizhuang
Joe Zhao. 2007. “JAK2V617F: Prevalence in a Large Chinese Hospital Population.” Blood 109 (1):
339–42. doi: 10.1182/blood-2006-03-009472.
Yamamoto, Yukiya, Hitoshi Kiyoi, Yasuyuki Nakano, Ritsuro Suzuki, Yoshihisa Kodera,
Shuichi Miyawaki, Norio Asou, et al. 2001. “Activating Mutation of D835 within the Activation
Loop of FLT3 in Human Hematologic Malignancies.” Blood 97 (8): 2434–9. doi: 10.1182/blood.
V97.8.2434.
Yan, Dongqing, Robert E. Hutchison, and Golam Mohi. 2012. “Critical Requirement for Stat5 in
a Mouse Model of Polycythemia Vera.” Blood 119 (15): 3539–49. doi: 10.1182/blood-2011-03­
345215.
Yano, Mio, Toshihiko Imamura, Daisuke Asai, Nobutaka Kiyokawa, Kazuhiko Nakabayashi,
Kenji Matsumoto, Takao Deguchi, et al. 2015. “Identification of Novel Kinase Fusion Transcripts
in Paediatric B Cell Precursor Acute Lymphoblastic Leukaemia with IKZF1 Deletion.” British
Journal of Haematology 171 (5): 813–17. doi: 10.1111/bjh.13757.
Yildiz, Mehmet, Li Hongxiu, Nisar A. Denzil Bernard, Peter Ouillette Amin, Siân Jones, Kamlai Saiya-
Cork, et al. 2015. “Activating STAT6 Mutations in Follicular Lymphoma.” Blood 125 (4): 668–79.
doi: 10.1182/blood-2014-06-582650.
Yoda, Akinori, Yuka Yoda, Sabina Chiaretti, Michal Bar-Natan, Kartik Mani, Scott J. Rodig,
Nathan West, et al. 2010. “Functional Screening Identifies CRLF2 in Precursor B-Cell Acute
Lymphoblastic Leukemia.” Proceedings of the National Academy of Sciences 107 (1): 252–7. doi:
10.1073/pnas.0911726107.
Younes, Anas, Jorge Romaguera, Michelle Fanale, Peter McLaughlin, Frederick Hagemeister,
Amanda Copeland, Sattva Neelapu, et al. 2012. “Phase I Study of a Novel Oral Janus Kinase 2
Inhibitor, SB1518, in Patients with Relapsed Lymphoma: Evidence of Clinical and Biologic Activity
in Multiple Lymphoma Subtypes.” Journal of Clinical Oncology 30 (33): 4161–7. doi: 10.1200/
JCO.2012.42.5223.
Zahn, Malena, Ralf Marienfeld, Ingo Melzner, Janine Heinrich, Benjamin Renner, Silke Wegener,
Anna Mießner, et al. 2017. “A Novel PTPN1 Splice Variant Upregulates JAK/STAT Activity in
Classical Hodgkin Lymphoma Cells.” Blood 129 (11): 1480–90. doi: 10.1182/blood-2016-06-720516.
Zeller, Bernward, Göran Gustafsson, Erik Forestier, Jonas Abrahamsson, Niels Clausen, Jesper Heldrup,
Liisa Hovi, et al. 2005. “Acute Leukaemia in Children with down Syndrome: A Population-Based
Nordic Study.” British Journal of Haematology 128 (6): 797–804. doi: 10.1111/j.1365­
2141.2005.05398.x.
JAK-STAT Signaling in Hematologic Malignancies 223

Zenatti, Priscila P., Daniel Ribeiro, Li Wenqing, Milene C. Linda Zuurbier, Maddalena Paganin Silva,
Julia Tritapoe, et al. 2011. “Oncogenic IL7R Gain-of-Function Mutations in Childhood T-Cell
Acute Lymphoblastic Leukemia.” Nature Genetics 43 (10): 932–9. doi: 10.1038/ng.924.
Zhao, Runxiang, Shu Xing, Li Zhe, Fu Xueqi, Li Qingshan, Sanford B. Krantz, and Zhizhuang Joe Zhao.
2005. “Identification of an Acquired JAK2 Mutation in Polycythemia Vera.” Journal of Biological
Chemistry 280 (24): 22788–92. doi: 10.1074/jbc.C500138200.
12
Aberrant JAK-STAT Signaling in Hematopoietic
Malignancies

Parvis Sadjadian
University Clinic for Hematology, Oncology, Hemostaseology and Palliative Care
Johannes Wesling Medical Center Minden, University of Bochum
Minden, Germany

12.1 Introduction
Many functions of the cells of the lympho-hematopoetic system like proliferation, differentiation,
and survival are regulated via cytokines and interferons (Commins et al. 2010). As most cytokines
lack intrinsic kinase activity, signaling is enabled using Janus kinases (JAK), a family of four
cytoplasmic non-receptor kinases (JAK1, JAK2, JAK3, TYK2). Downstream, JAK phosphorylates
cytoplasmic proteins called STAT (signal transducers and activators of transcription), a group of
seven (STAT1, STAT2, STAT3, STAT4, STAT5A, STAT5B and STAT6) transcription factors,
which after activation dimerize and translocate to the nucleus to bind DNA for regulation of gene
expression and initiation of transcription (Figure 12.1). The JAK-STAT pathway is an extremely
fast membrane-to-nucleus signaling system, whereby extracellular factors control gene expression.
Interestingly, STAT activity in the nucleus can be detected only minutes after cytokine binding
(Leonard and O’Shea 1998). JAK-mediated signaling is essential for both short-chain cytokines
(like IL-2, IL-3, IL-4, IL-5, GM-CSF, IL-7, IL-9, IL-13, IL-15) and long-chain cytokines (like IL-6,
IL-11, OSM, CNTF, CT-1, growth hormone, prolactin, erythropoietin, thrombopoietin) (Leonard
and O’Shea 1998). Distinct cytokines and their associated receptors use selective JAKs and STATs
for signaling, explaining their different in vivo roles (O’Shea et al. 2015).
More than 25 years ago, the JAK-STAT pathway was first described investigating interferon
signaling (Velazquez et al. 1992, Darnell et al. 1994). Enabling cytokine-regulated cell proliferation,
differentiation and survival, it soon became clear that this pathway has a critical role in tumorigenesis
(Bromberg et al. 1999, Levy and Darnell 2002), both in solid tumors and hematologic malignancies.
The JAK-STAT pathway can be pathologically activated by several mechanisms like autocrine/
paracrine cytokine production, mutations in upstream oncogenes (like BCR-ABL in CML or FLT3
in AML), or rare activating mutations of STATs. This overview will primarily focus on mutations of
JAKs leading to various hematopoetic malignancies. In addition, therapeutic options to target the
JAK-STAT pathway will be elucidated, with both experimental approaches and therapies already
implemented in clinical routine.

12.2 Dysregulation of the JAK-STAT Pathway in Hematological Malignancies


Alterations of the JAK-STAT pathway are of relevance in hematopoietic malignancies (Benekli
et al. 2003). For example, oncogenic JAK2 rearrangements with multiple fusion gene partners have
been reported in a variety of hematologic malignancies (Hammaren et al. 2019). However, these

225
226 JAK-STAT Signaling in Diseases

FIGURE 12.1 Illustration of the role of driver mutations in the pathogenesis of MPN. Left: Activation mutation in
the thrombopoetin receptor. Middle: JAK2 (V617F or exon12) gain of function mutation. Right: Calreticulin mutation
(TPO: thrombopoietin; JAK2: Janus kinase 2; CALR: calreticulin; MPL: myeloproliferative leukemia virus oncogene;
STAT: signal transducer and activator of transcription).

translocations are rare: the Mayo Clinic cytogenetics database spanning the years 1989–2008 with
24,262 patients showed JAK2 translocations in only 0.06% of hematopoietic neoplasms (Patnaik
et al. 2010). In contrast, JAK2 V617F mutations in exon 14 of JAK2 is prevalent in 50–60% of
patients with MPNs (in polycythemia vera (PV) even 95%), as compared to 0.1% in the general
population (Xu et al. 2007, Nielsen et al. 2014).
In the following sections, the impact of JAK-STAT alterations on the pathogenesis, course, and
prognosis of major myeloid and lymphoid malignancies are discussed.

12.2.1 Myeloid Malignancies


12.2.1.1 Myeloproliferative Neoplasms
The classical BCR-ABL-negative MPN essential thrombocythemia (ET), PV, and primary
myelofibrosis (PMF) are closely related stem-cell-derived disorders characterized by increased
proliferation of erythroid, megakaryocytic, or granulocytic cells in the absence of dysplasia
(Dameshek 1951). The genetic basis of relationship was discovered in 2005 with the identifica­
tion of the acquired mutation in exon 14 of JAK2 (V617F, a substitution of valine to
phenylalanine at codon 617), which can be found in almost all patients with PV and more
than 50% of patients with ET or PMF (Baxter et al. 2005, James et al. 2005, Kralovics et al.
2005, Levine et al. 2005b). JAK2 V617F is a gain-of-function mutation in the JH2 (JAK
JAK-STAT in Hematopoietic Malignancies 227

homology 2) pseudokinase domain, a region that inhibits JAK2 kinase activity (Saharinen and
Silvennoinen 2002, Saharinen et al. 2003).
Subsequently, other driver mutations in calreticulin (CALR, an endoplasmic reticulin chaper­
one) or myeloproliferative leukemia virus oncogene (MPL, the thrombopoietin receptor) were
identified (Chaligne et al. 2007, Klampfl et al. 2013, Nangalia et al. 2013, Pardanani et al. 2006).
All lead to a deregulated JAK-STAT pathway via increased JAK2 signaling (Figure 12.1),
resulting in abnormally high circulating levels of pro-inflammatory cytokines (Rampal et al.
2014, Skoda et al. 2015).
Within the three classical MPNs, the frequency of driver mutations is varying (Figure 12.2)
(Griesshammer and Sadjadian 2017). Interestingly, even in rare MPN cases not carrying one of the
three driver mutations (approximately 10% of PMF and 15% of ET cases), there is an overactive
JAK-STAT pathway, too (Skoda et al. 2015). Thus, JAK-inhibiting strategies emerged as
a promising concept of treating MPN soon after the discovery of the JAK2 V617F point mutation
in 2005 (Griesshammer and Sadjadian 2017).

12.2.1.2 Acute Myeloid Leukemia


Acute leukemias—both acute myeloid leukemia (AML) and acute lymphoblastic leukemia (ALL)—
are characterized by a differentiation block causing the accumulation of immature blast cells in the
bone marrow and/or peripheral blood.
Disturbances of the JAK-STAT pathway are uncommon in AML. In a larger series of
genotypic analyses of patients with hematopoietic malignancies only four of 222 patients
(1.8%) with AML were positive for JAK2 V617F mutations, three of whom had been
previously diagnosed with MPN (Levine et al. 2005b). A group from Cambridge reported
JAK2 V617F mutations in five of 90 (4%) primary hematopoietic cells (Scott et al. 2005).
Steesma et al. found JAK2 V617F mutations in 13 of 162 AML samples (8%), but 10 of 13
were secondary post MPN AML patients and only three were de novo AML patients
(Steensma et al. 2006).
Several analogous JAK2 gene fusions have been described in AML (Murati et al. 2005, Reiter
et al. 2005). The mechanism of JAK2 activation is assumed to be similar in all cases, with the
JAK2 fusion partner promoting dimerization and constitutive activation of the JAK2 tyrosine
kinase (Chen et al. 2012).
Interestingly, in a study of 77 AML patients, the detection of high levels of activated JAK2 (by
immunohistochemistry using a phosphor specific antibody against JAK2) predicted a poor outcome
(Ikezoe et al. 2011).

FIGURE 12.2 Frequencies of driver mutations JAK2 exon12 and exon14, CALR and MPL in ET, PV and PMF. In
PMF and ET about 10–15% do not have any driver mutation (so-called triple negative), in PV around 2% do not carry
a JAK2 (neither exon 12 nor exon 14) mutation.
228 JAK-STAT Signaling in Diseases

12.2.1.3 Myelodysplastic Syndrome


Myelodysplastic syndromes (MDS) are a group of clonal disorders of the hematopoietic stem cell
characterized by dysplasia of bone marrow and peripheral blood cells, hematopoietic insufficiency
(mostly anemia), and an elevated risk of transformation to AML.
JAK2 mutations also rare in MDS, only two of 48 patients (4.2%) with MDS were tested
positive for JAK2 V617F mutation (Levine et al. 2005b). Interestingly, in MDS/MPN with ring
sideroblasts and thrombocytosis (formerly: refractory anemia with ring sideroblasts and thrombo­
cytosis, RARS-T), a myeloid neoplasm showing clinical and morphological features that overlap
between MDS and MPN [study from Arber et al. 2016], JAK2 V617F mutations could be found
in 54 of 93 patients (58%) (Wardrop and Steensma 2009, Arber et al. 2016).

12.2.1.4 Chronic Myelomonocytic Leukemia


Chronic myelomonocytic leukemia (CMML), one of the MDS/MPN overlap syndromes, is char­
acterized by peripheral blood monocytosis, bone marrow dysplasia with transfusion-dependent
cytopenias, and an elevated risk of progression to AML. Typically, CMML is dependent on aberrant
signaling through the granulocyte-macrophage colony-stimulating factor pathway, which requires
JAK-STAT (Padron et al. 2013, Bose and Verstovsek 2017). Acknowledging the myeloproliferative
component of the disease makes JAK-inhibition a reasonable approach (Solary 2016). A phase I trial
of ruxolitinib in CMML showed encouraging results with hematological or spleen responses in seven
of 20 patients (35%) (Padron et al. 2016).

12.2.2 Lymphoid Malignancies


Activation of the JAK-STAT pathway by cytokines such as IL-2, IL-4, IL-7, IL-9, IL-15, and IL-21
is essential in the formation and function of lymphoid cells like T-cells, B-cells, and NK-cells (Liao
et al. 2011). Receptor signaling (JAK1 via the common cytokine-binding chains, JAK3 via the
common γ-chain) activates typically STAT3 and STAT5 (Vainchenker and Constantinescu 2013).

12.2.2.1 Acute Lymphoblastic Leukemia


Acute lymphoblastic leukemia was the first malignancy that was associated with JAK-signaling
pathway disturbances. In 1997, Lacronique et al. discovered a TEL-JAK2 fusion gene as
a product of a t(9;12)(p24;p13) translocation in a patient with a T-ALL (Lacronique et al. 1997).
Subsequently, other JAK2 fusion proteins with different fusion partners like PCM1-JAK2, PAX5­
JAK2 or SSBP2-JAK2 were discovered in ALL, all with the same mechanism of JAK2-activation
via facilitating dimerization and constitutive activation of the JAK2 tyrosine kinase component
(Reiter et al. 2005, Poitras et al. 2008, Nebral et al. 2009).
Sequencing of exon14 of the JAK2 gene in patients with Down syndrome-associated B-progenitor
ALL revealed acquired somatic JAK2 mutations (R683G being the most common) in 18–28%, also
in 7% of high-risk pediatric B-ALL patients irrespective of having Down syndrome (Bercovich et al.
2008, Kearney et al. 2009, Mullighan et al. 2009).
JAK3 mutations in ALL are rare; however, JAK1 mutations have been reported in 4–18% of
patients with T-ALL (Flex et al. 2008, Asnafi et al. 2010).

12.2.2.2 B-Cell Non-Hodgkin Lymphoma


Non-Hodgkin lymphomas (NHL) such as diffuse large B-cell lymphoma (DLBCL), follicular
lymphoma, marginal zone lymphoma, and mantle cell lymphoma are a heterogeneous group of
malignant lymphoproliferative disorders predominantly originating in B-lymphocytes.
In activated B-cell-like diffuse large B-cell lymphoma (ABC-DLBCL), one of the two pheno­
typic subtypes, high-level STAT3 expression and activation can be detected (Ding et al. 2008).
JAK-STAT in Hematopoietic Malignancies 229

Activation of STAT3 and its kinase, JAK1, is caused by autocrine production of IL-6 and IL-10.
STAT3 regulates multiple oncogenic signaling pathways, including NF-κB, a cell-cycle checkpoint,
PI3K/AKT/mTORC1, and STAT3 itself. In addition, STAT3 negatively regulates the lethal type
I IFN signaling pathway (Lu et al. 2018). In DLBCL, STAT3 activity is associated with a poor
prognosis (Huang et al. 2013, Ok et al. 2014).

12.2.2.3 Chronic Lymphocytic Leukemia


Chronic lymphocytic leukemia (CLL) is an indolent B-cell non-Hodgkin lymphoma characterized by
a leukemic course of disease, representing the most common form of adult leukemia in western countries.
The JAK2 V617F-mutation is not present in CLL, but constitutive phosphorylation and
activation of STATs has been described (Frank et al. 1997, Levine et al. 2005a).

12.2.2.4 Hodgkin Lymphoma


Hodgkin lymphoma is a rare malignant disease involving lymph nodes and the lymphatic system.
Biopsies typically show a limited number of the malignant Reed-Sternberg cells surrounded by
numerous reactive bystander cells.
The JAK/STAT pathway is an essential regulator of Hodgkin lymphoma cell viability, and
JAK-mediated signaling has been demonstrated in a significant number of patients (Baus and
Pfitzner 2006, Navarro et al. 2009). In Hodgkin’s lymphoma, JAK-STAT is primarily activated by
an aberrant deregulation of a network of cytokines and chemokines in the microenvironment
(Derenzini and Younes 2013).
JAK2 amplifications and gain-of-function mutations have been observed in up to a third of
patients with Hodgkin’s lymphoma. The same alterations were found in 30–50% of patients with
primary mediastinal B-cell lymphoma (Joos et al. 2000, Rosenwald et al. 2003, Green et al. 2010).
JAK2 rearrangements occur in a minority of classical Hodgkin lymphoma cases; screening of
131 patients identified one case with a SEC31A-JAK2 rearrangement and two additional cases
with rearrangements involving JAK2 (Van Roosbroeck et al. 2011).

12.2.2.5 Multiple Myeloma


Multiple myeloma is a heterogeneous hematologic neoplasia characterized by monoclonal proliferation
and accumulation of malignant plasma cells in the bone marrow leading to bone marrow insufficiency,
bone destruction, and increased production of complete and incomplete immunoglobulins.
The therapeutic armamentarium in this disease, which accounts for approximately 10% of all
hematologic malignancies, has dramatically improved over the past years. However, most of these
patients eventually relapse. One of the pathways involved in the pathogenesis is the JAK-STAT3
axis via STAT3 activation by several cytokines including IL-6, which can be secreted by bone
marrow stromal cells or myeloma cells themselves (Mitsiades et al. 2006, Podar et al. 2009,
Sansone and Bromberg 2012, Hu and Hu 2018). Preclinical experiments demonstrated decreased
survival of primary myeloma cells or myeloma cell lines using a STAT3 inhibitor (Nelson et al.
2008). Combining the JAK1/2 inhibitor ruxolitinib with the proteasome inhibitor bortezomib, one
of the standard drugs in myeloma treatment, showed a decrease of many JAK-STAT pathway
genes (de Oliveira et al. 2017).

12.2.3 Non-Malignant Hematologic Disorders: Immune Thrombocytopenic Purpura


(Idiopathic Thrombocytopenia, ITP)
ITP is an acquired form of thrombocytopenia induced by an autoimmune reaction targeting
thrombocytes and megakaryocytes. The underlying cause is mostly unknown and secondary
forms (e.g. drug-related) account for approximately 20% (Moulis et al. 2014).
230 JAK-STAT Signaling in Diseases

A recent publication from China elucidated the role of the JAK-STAT pathway in the
pathogenesis of immune thrombocytopenia (ITP) (Zhang et al. 2018). In total, 26 patients with
confirmed ITP were compared with 24 healthy controls concerning the expression of JAK3,
pJAK3 mRNA, STAT3, and pSTAT3. In addition, changes in levels of IL21 mRNA, IL-21 cell
secretion after dexamethasone intervention and AG490 (a specific and potent inhibitor of JAK2)
blockade were measured. Compared with the healthy controls, patients with ITP had
a significantly higher expression of JAK3, pJAK3 mRNA, STAT3, and pSTAT3 protein, which
were significantly lower after blocking of AG490. The expression of IL-21 mRNA and the
secretion of IL-21 decreased after Dexamethasone intervention, but significantly increased after
AG490 blocking. These results suggest that the IL-21 activated JAK-STAT signaling pathway
may play a role both in the pathogenesis and in the regulation of ITP (Zhang et al. 2018).
Therefore, activation of the JAK-STAT pathway could be used to treat ITP (Evangelista et al.
2007, Garnock-Jones and Keam 2009, Zhang and Kolesar 2011, Kuter 2013): eltrombopag, that
activates the JAK-STAT pathway as a thrombopoietin (TPO) non-peptide mimetic, has been
evaluated in clinical use and was approved by the FDA (2008) and EMA (2010) to treat patients
with ITP who have had an insufficient response to corticosteroids or immunoglobulin therapy.

12.3 Therapeutic Implications


12.3.1 JAK Inhibitors in the Treatment of Hematopoietic Malignancies
12.3.1.1 Ruxolitinib
Ruxolitinib (INCB018424/INC424) is a potent and selective oral inhibitor of JAK 1 and JAK2.
Currently, Ruxolitinib is licensed for the treatment of myelofibrosis by the US Food and Drug
Administration (FDA) since 2011 and by the European Medicines Agency (EMA) since 2012,
followed by the approval for the treatment of hydroxyurea-resistant or -intolerant PV in 2014.
Approval of ruxolitinib for myelofibrosis was based on the two pivotal phase III trials
COMFORT I (ruxolitinib vs. placebo) and COMFORT-II (ruxolitinib vs. best available therapy)
(Harrison et al. 2012, Verstovsek et al. 2012). In both COMFORT trials, crossover was allowed.
Ruxolitinib was superior in reducing spleen size, improving myelofibrosis-associated symptoms
and quality of life. The most frequent hematological side effects of ruxolitinib treatment were
dose-dependent anemia and thrombocytopenia. Long-term follow-up demonstrated an on-going
preserved treatment effect of ruxolitinib (Harrison et al. 2016). In pooled analyses of COMFORT­
1 and COMFORT-2, a prolonged overall survival of ruxolitinib treated patients was suggested
compared with controls (Verstovsek et al. 2017).
In PV, ruxolitinib was licensed based on the RESPONSE-1 and RESPONSE-2 phase III trials
(Vannucchi 2015, Verstovsek et al. 2016, Passamonti et al. 2017, Griesshammer et al. 2018). In PV
patients who had an inadequate response to or unacceptable side effects from previous hydro­
xyurea treatment, ruxolitinib was superior at controlling hematocrit levels and the number of
phlebotomy-free patients was doubled. Additionally, ruxolitinib led to an improvement both in
symptom burden and quality of life (Passamonti et al. 2017).
Ruxolitinib has been evaluated in various hematologic indications, albeit clinical data are
limited. In a phase II trial, 38 patients with refractory leukemia (12 with JAK2 V617F mutation)
received ruxolitinib (Eghtedar et al. 2012). The anti-leukemic potential of ruxolitinib was modest,
and three patients showed a significant response (all of which with a secondary post-MPN AML).
As there is a strong rationale for targeting JAK-STAT signaling, ruxolitinib was evaluated in
relapsed/refractory Hodgkin lymphoma (Baus and Pfitzner 2006, Lee et al. 2018b). In a recent phase
II trial, ruxolitinib demonstrated a rapid B-symptom relief and a hematological response in six of 32
patients (18.8%), although lasting only little more than 7 months (Van Den Neste et al. 2018). Now,
ruxolitinib is evaluated in various non-MPN indications in several clinical trials (see Table 12.1).
JAK-STAT in Hematopoietic Malignancies 231

TABLE 12.1
Interventional Clinical Trials Evaluating Ruxolitinib in the Setting of Leukemia or Lymphoma (https://
clinicaltrials.gov, accessed November 1, 2018)
Identifier Indication Phase Intervention Status

NCT03722407 CMML 2 Ruxolitinib Not yet recruiting


NCT03681561 r/r classical Hodgkin lymphoma 1/2 Ruxolitinib/nivolumab Recruiting
NCT03571321 ph-like ALL (young adults) 1 Ruxolitinib Not yet recruiting
NCT03110822 r/r multiple myeloma 1 Ruxolitinib/lenalidomide Recruiting
NCT02974647 T or NK cell lymphoma 2 Ruxolitinib Recruiting
NCT02912754 CLL 1/2 Ruxolitinib/ibrutinib Not yet recruiting
NCT02723994 pediatric ALL 2 Ruxolitinib Recruiting
NCT02613598 r/r lymphoma 1 Ruxolitinib/bortezomid Recruiting
NCT02257138 r/r or post-MPN AML 1/2 Ruxolitinib/decitabine Recruiting
NCT02164500 classical Hodgkin lymphoma 2 Ruxolitinib Recruiting
NCT02131584 CLL 2 Ruxolitinib Active, not recruiting
NCT02015208 CLL 1/2 Ruxolitinib Completed
NCT01776723 CMML 1/2 Ruxolitinib Active, not recruiting
NCT01712659 T-cell leukemia (adults) 2 Ruxolitinib Recruiting
NCT01431209 r/r DLBCL or T-cell-NHL 2 Ruxolitinib Active, not recruiting
NCT00674479 AML, ALL, CML, MDS 2 Ruxolitinib Completed

Ruxolitinib like most other JAK inhibitors in clinical evaluation is a type I kinase inhibitor that
binds in the ATP binding pocket of the kinase domain in its active configuration (Liu and Gray
2006). It ameliorates splenomegaly and improves constitutional symptom burden, but has only
a modest effect on the mutant allele burden and has a limited disease modifying potential (Harrison
et al. 2012, Verstovsek et al. 2012, Vannucchi et al. 2017). However, a recently published trial in
myelofibrosis patients receiving long-term ruxolitinib demonstrated a stabilization or improvement
of fibrosis versus a control cohort receiving best available therapy (Kvasnicka et al. 2018). Type II
JAK inhibitors like CHZ868, which preferentially inhibit V617F mutant versus wildtype JAK2 thus
stabilizing JAK2 in the inactive conformation, showed increased efficacy in preclinical MPN models
and reduced mutant allele burden in JAK2-/MPL mutant MPN models (Meyer et al. 2015,
Silvennoinen and Hubbard 2015).

12.3.1.2 Other JAK Inhibitors


Dose escalation of ruxolitinib is limited due to myelotoxicity, excluding many patients with baseline
cytopenias from treatment. Novel JAK inhibitors with less potential to induce thrombocytopenia
and even properties to improve anemia could offer additional treatment options for a number of
patients, especially those with baseline cytopenias (Griesshammer and Sadjadian 2017).
Pacritinib (SB-1518) is a small molecular weight macrocyclic oral JAK2-selective inhibitor
(JAK2 wild type and V617F mutant) that also targets FLT3 (FMS-like tyrosine kinase 3, an
important target in the treatment of AML), CSF1R and IRAK1, and no meaningful inhibition of
JAK1 (Poulsen et al. 2012, Hatzimichael et al. 2014). Pacritinib was evaluated in two recent phase
III trials (PERSIST-1 and PERSIST-2) in patients with myelofibrosis regardless of thrombocyte
count and in patients with thrombocytopenic myelofibrosis (Mesa et al. 2017b, Mascarenhas et al.
2018). Compared to best available therapy, pacritinib met the primary end point of superior
spleen volume reduction, but failed to show a significant improvement in total symptom score
which was the co-primary endpoint.
232 JAK-STAT Signaling in Diseases

Pacritinib has encouraging activity in lymphoma. In a phase I trial of 34 patients with relapsed/
refractory Hodgkin or non-Hodgkin lymphoma, three patients had a partial remission and 15
patients had a stable disease (Younes et al. 2012).
Pacritinib could also be an interesting option in the therapy of AML as up to a third of AML
patients harbor FLT3 mutations. An activation of the JAK-STAT pathway has been observed
after initiation of FLT3 tyrosine kinase inhibitor therapy (Hart et al. 2011). Thus, pacritinib could
reduce the development of secondary resistance or serve as a second-line therapy to re-sensitize
resistant cells to FLT3 inhibition (Weisberg et al. 2012). Clinical results with the potential to
change treatment algorithms are eagerly awaited.
Another important drug in advanced clinical development is momelotinib (CYT387), a dual
JAK1/2 inhibitor (Pardanani et al. 2009, Tyner et al. 2010). Momelotinib demonstrated a unique
property of improving myelofibrosis-associated anemia (Pardanani et al. 2013). However, in two
phase III trials in myelofibrosis (SIMPLIFY-1 and SIMPLIFY-2), momelotinib was not superior
to best available therapy concerning spleen size reduction, but it significantly improved disease-
related symptoms and reduced transfusion dependency (Mesa et al. 2017a, Harrison et al. 2018).

12.3.2 STAT Inhibitors


Blocking the JAK-STAT pathway with STAT inhibitors is an intriguing approach for anticancer
therapy. Limitations are the selectivity, bioavailability, similar homology (e.g. STAT3 and STAT1),
and poor in vivo efficacy (O’Shea et al. 2015).
Several STAT inhibitors are in clinical development; however, due to the problems mentioned
above none has been successfully introduced into clinical routine (Mandal et al. 2011). Different
strategies may be used to design inhibitors targeting STAT: e.g. blocking of phosphorylation,
preventing of dimerization by blocking the SH2 domain, or interfering with DNA-binding (Seif
et al. 2017, Lee et al. 2018a).
Many publications and studies have demonstrated the important role of STAT (especially
STAT3) in numerous aspects of tumorigenesis like differentiation, proliferation, angiogenesis,
apoptosis, and metastasis (Jove 2000, Levy and Inghirami 2006, Yu et al. 2009). Recently,
a study from Taiwan showed that JAK-STAT3 signaling is the major pathway in DLBCL
(diffuse large B-cell lymphoma) amoeboid movement, which is critical for early dissemination
of the disease (Pan et al. 2018). STAT3 is activated in response to cytokines (like IL-6, IL-10,
IL-21, IL-27, leptin or G-CSF) or receptor and non-receptor tyrosine-kinases (Sansone and
Bromberg 2012).
AZD9150 (ISIS481464) is an antisense oligonucleotide inhibitor of STAT3. At ASCO 2013,
Hong and colleagues presented a phase I study in metastatic cancer that showed promising results
using AZD9150 as a single agent. In addition, two out of three patients with chemotherapy-
refractory diffuse large B-cell lymphoma had prolonged partial responses ( Seth et al. 2010, Hong
et al. 2015). As STAT3 blocking rather slows tumor progression than having antineoplastic
effects, AZD9150 is now evaluated in several clinical trials in combination with checkpoint
inhibitors (Kroemer et al. 2016, Pianko et al. 2017, Syed 2017, Cullberg et al. 2018).
Moreover, in vitro tests with existing drugs inhibiting STAT5 like the psychotropic drug
pimozide have shown that STAT5 inhibitors may be effective in improving treatment outcome in
acute or chronic myelogenous leukemia (Nelson et al. 2011, 2012).

12.4 Summary
The JAK-STAT signaling pathway is playing a central role in the formation of blood cells via
cytokine activation of transmembrane receptors with transmission of anti-apoptotic, proliferative,
or differentiation signals.
JAK-STAT in Hematopoietic Malignancies 233

Several hematopoietic malignancies have been associated with mutations of JAK, cytokine
receptors, or other components of the JAK-STAT signaling pathway. All these alterations are able
to drive the disease. Today, most data are available in MPNs, where the presence of mutations in
JAK2, calreticulin, or MPL has also been implemented as major diagnostic criteria (Arber et al.
2016). Moreover, ruxolitinib is currently the only licensed drug successfully targeting the JAK­
STAT pathway in MPN and is now a milestone in the armamentarium of MPN therapy.
STAT inhibitors, particularly targeting STAT3, are now evaluated in many hematopoietic
malignancies and could potentially emerge as valuable treatment options in the future.

REFERENCES
Arber, D. A., A. Orazi, R. Hasserjian, J. Thiele, M. J. Borowitz, M. M. Le Beau, C. D. Bloomfield,
M. Cazzola, and J. W. Vardiman. 2016. “The 2016 revision to the World Health Organization
classification of myeloid neoplasms and acute leukemia.” Blood 127 (20): 2391–2405. doi: 10.1182/
blood-2016-03-643544.
Asnafi, V., S. Le Noir, L. Lhermitte, C. Gardin, F. Legrand, X. Vallantin, J. V. Malfuson, N. Ifrah,
H. Dombret, and E. Macintyre. 2010. “JAK1 mutations are not frequent events in adult T-ALL:
a GRAALL study.” Br J Haematol 148 (1): 178–179. doi: 10.1111/j.1365-2141.2009.07912.x.
Baus, D., and E. Pfitzner. 2006. “Specific function of STAT3, SOCS1, and SOCS3 in the regulation of
proliferation and survival of classical Hodgkin lymphoma cells.” Int J Cancer 118 (6):1404–1413.
doi: 10.1002/ijc.21539.
Baxter, E. J., L. M. Scott, P. J. Campbell, C. East, N. Fourouclas, S. Swanton, G. S. Vassiliou, A. J. Bench,
E. M. Boyd, N. Curtin, M. A. Scott, W. N. Erber, A. R. Green, and Project Cancer Genome. 2005.
“Acquired mutation of the tyrosine kinase JAK2 in human myeloproliferative disorders.” Lancet
365 (9464):1054–1061. doi: 10.1016/S0140-6736(05)71142-9.
Benekli, M., M. R. Baer, H. Baumann, and M. Wetzler. 2003. “Signal transducer and activator of
transcription proteins in leukemias.” Blood 101 (8):2940–2954. doi: 10.1182/blood-2002-04-1204.
Bercovich, D., I. Ganmore, L. M. Scott, G. Wainreb, Y. Birger, A. Elimelech, C. Shochat, G. Cazzaniga,
A. Biondi, G. Basso, G. Cario, M. Schrappe, M. Stanulla, S. Strehl, O. A. Haas, G. Mann, V. Binder,
A. Borkhardt, H. Kempski, J. Trka, B. Bielorei, S. Avigad, B. Stark, O. Smith, N. Dastugue,
J. P. Bourquin, N. B. Tal, A. R. Green, and S. Izraeli. 2008. “Mutations of JAK2 in acute
lymphoblastic leukaemias associated with Down’s syndrome.” Lancet 372 (9648): 1484–1492. doi:
10.1016/S0140-6736(08)61341-0.
Bose, P., and S. Verstovsek. 2017. “JAK2 inhibitors for myeloproliferative neoplasms: what is next?”
Blood 130 (2): 115–125. doi: 10.1182/blood-2017-04-742288.
Bromberg, J. F., M. H. Wrzeszczynska, G. Devgan, Y. Zhao, R. G. Pestell, C. Albanese, and J. E. Darnell,
Jr. 1999. “Stat3 as an oncogene.” Cell 98 (3):295–303. doi: 10.1016/s0092-8674(00)81959-5.
Chaligne, R., C. James, C. Tonetti, R. Besancenot, J. P. Le Couedic, F. Fava, F. Mazurier, I. Godin,
K. Maloum, F. Larbret, Y. Lecluse, W. Vainchenker, and S. Giraudier. 2007. “Evidence for MPL
W515L/K mutations in hematopoietic stem cells in primitive myelofibrosis.” Blood 110 (10):3735–
3743. doi: 10.1182/blood-2007-05-089003.
Chen, E., L. M. Staudt, and A. R. Green. 2012. “Janus kinase deregulation in leukemia and lymphoma.”
Immunity 36 (4): 529–541. doi: 10.1016/j.immuni.2012.03.017.
Commins, S. P., L. Borish, and J. W. Steinke. 2010. “Immunologic messenger molecules: cytokines,
interferons, and chemokines.” J Allergy Clin Immunol 125 (2 Suppl 2): S53–S72. doi: 10.1016/j.
jaci.2009.07.008.
Cullberg, M., C. Arfvidsson, B. Larsson, A. Malmgren, P. Mitchell, U. Wahlby Hamren, and H. Wray.
2018. “Pharmacokinetics of the Oral Selective CXCR2 Antagonist AZD5069: a summary of eight
phase I studies in healthy volunteers.” Drugs R D 18 (2): 149–159. doi: 10.1007/s40268-018-0236-x.
Dameshek, W. 1951. “Some speculations on the myeloproliferative syndromes.” Blood 6 (4): 372–375.
Darnell, J. E., Jr., I. M. Kerr, and G. R. Stark. 1994. “Jak-STAT pathways and transcriptional activation
in response to IFNs and other extracellular signaling proteins.” Science 264 (5164): 1415–1421. doi:
10.1126/science.8197455.
234 JAK-STAT Signaling in Diseases

de Oliveira, M. B., V. L. Fook-Alves, A. I. P. Eugenio, R. C. Fernando, L. F. G. Sanson, M. F. de


Carvalho, W. M. T. Braga, F. E. Davies, and G. W. B. Colleoni. 2017. “Anti-myeloma effects of
ruxolitinib combined with bortezomib and lenalidomide: a rationale for JAK/STAT pathway
inhibition in myeloma patients.” Cancer Lett 403: 206–215. doi: 10.1016/j.canlet.2017.06.016.
Derenzini, E., and A. Younes. 2013. “Targeting the JAK-STAT pathway in lymphoma: a focus on
pacritinib.” Expert Opin Investig Drugs 22 (6):775–785. doi: 10.1517/13543784.2013.775244.
Ding, B. B., J. J. Yu, R. Y. Yu, L. M. Mendez, R. Shaknovich, Y. Zhang, G. Cattoretti, and B. H. Ye. 2008.
“Constitutively activated STAT3 promotes cell proliferation and survival in the activated B-cell subtype
of diffuse large B-cell lymphomas.” Blood 111 (3): 1515–1523. doi: 10.1182/blood-2007-04-087734.
Eghtedar, A., S. Verstovsek, Z. Estrov, J. Burger, J. Cortes, C. Bivins, S. Faderl, A. Ferrajoli,
G. Borthakur, S. George, P. A. Scherle, R. C. Newton, H. M. Kantarjian, and F. Ravandi. 2012.
“Phase 2 study of the JAK kinase inhibitor ruxolitinib in patients with refractory leukemias,
including postmyeloproliferative neoplasm acute myeloid leukemia.” Blood 119 (20): 4614–4618.
doi: 10.1182/blood-2011-12-400051.
Evangelista, M. L., S. Amadori, and R. Stasi. 2007. “Biologic aspects of thrombopoietin and the
development of novel thrombopoietic agents for clinical use.” Curr Drug Discov Technol 4 (3):
162–173.
Flex, E., V. Petrangeli, L. Stella, S. Chiaretti, T. Hornakova, L. Knoops, C. Ariola, V. Fodale, E. Clappier,
F. Paoloni, S. Martinelli, A. Fragale, M. Sanchez, S. Tavolaro, M. Messina, G. Cazzaniga, A. Camera,
G. Pizzolo, A. Tornesello, M. Vignetti, A. Battistini, H. Cave, B. D. Gelb, J. C. Renauld, A. Biondi,
S. N. Constantinescu, R. Foa, and M. Tartaglia. 2008. “Somatically acquired JAK1 mutations in
adult acute lymphoblastic leukemia.” J Exp Med 205 (4): 751–758. doi: 10.1084/jem.20072182.
Frank, D. A., S. Mahajan, and J. Ritz. 1997. “B lymphocytes from patients with chronic lymphocytic
leukemia contain signal transducer and activator of transcription (STAT) 1 and STAT3 constitutively
phosphorylated on serine residues.” J Clin Invest 100 (12): 3140–3148. doi: 10.1172/JCI119869.
Garnock-Jones, K. P., and S. J. Keam. 2009. “Eltrombopag.” Drugs 69 (5): 567–576. doi: 10.2165/
00003495-200969050-00005.
Green, M. R., S. Monti, S. J. Rodig, P. Juszczynski, T. Currie, E. O’Donnell, B. Chapuy, K. Takeyama,
D. Neuberg, T. R. Golub, J. L. Kutok, and M. A. Ship. 2010. “Integrative analysis reveals selective
9p24.1 amplification, increased PD-1 ligand expression, and further induction via JAK2 in nodular
sclerosing Hodgkin lymphoma and primary mediastinal large B-cell lymphoma.” Blood 116 (17):
3268–3277. doi: 10.1182/blood-2010-05-282780.
Griesshammer, M., and P. Sadjadian. 2017. “The BCR-ABL1-negative myeloproliferative neoplasms:
a review of JAK inhibitors in the therapeutic armamentarium.” Expert Opin Pharmacother 18 (18):
1929–1938. doi: 10.1080/14656566.2017.1404574.
Griesshammer, M., G. Saydam, F. Palandri, G. Benevolo, M. Egyed, J. Callum, T. Devos, S. Sivgin,
P. Guglielmelli, C. Bensasson, M. Khan, J. P. Ronco, and F. Passamonti. 2018. “Ruxolitinib for the
treatment of inadequately controlled polycythemia vera without splenomegaly: 80-week follow-up
from the RESPONSE-2 trial.” Ann Hematol 97 (9): 1591–1600. doi: 10.1007/s00277-018-3365-y.
Hammaren, H. M., A. T. Virtanen, J. Raivola, and O. Silvennoinen. 2019. “The regulation of JAKs in
cytokine signaling and its breakdown in disease.” Cytokine 118: 48–63. doi: 10.1016/j.
cyto.2018.03.041.
Harrison, C., J. J. Kiladjian, H. K. Al-Ali, H. Gisslinger, R. Waltzman, V. Stalbovskaya, M. McQuitty,
D. S. Hunter, R. Levy, L. Knoops, F. Cervantes, A. M. Vannucchi, T. Barbui, and G. Barosi. 2012.
“JAK inhibition with ruxolitinib versus best available therapy for myelofibrosis.” N Engl J Med 366
(9): 787–798. doi: 10.1056/NEJMoa1110556.
Harrison, C. N., A. M. Vannucchi, J. J. Kiladjian, H. K. Al-Ali, H. Gisslinger, L. Knoops, F. Cervantes,
M. M. Jones, K. Sun, M. McQuitty, V. Stalbovskaya, P. Gopalakrishna, and T. Barbui. 2016.
“Long-term findings from COMFORT-II, a phase 3 study of ruxolitinib vs best available therapy
for myelofibrosis.” Leukemia 30 (8): 1701–1707. doi: 10.1038/leu.2016.148.
Harrison, C. N., A. M. Vannucchi, U. Platzbecker, F. Cervantes, V. Gupta, D. Lavie, F. Passamonti,
E. F. Winton, H. Dong, J. Kawashima, J. D. Maltzman, J. J. Kiladjian, and S. Verstovsek. 2018.
“Momelotinib versus best available therapy in patients with myelofibrosis previously treated with
ruxolitinib (SIMPLIFY 2): a randomised, open-label, phase 3 trial.” Lancet Haematol 5 (2): e73–e81.
doi: 10.1016/S2352-3026(17)30237-5.
JAK-STAT in Hematopoietic Malignancies 235

Hart, S., K. C. Goh, V. Novotny-Diermayr, Y. C. Tan, B. Madan, C. Amalini, L. C. Ong, B. Kheng,


A. Cheong, J. Zhou, W. J. Chng, and J. M. Wood. 2011. “Pacritinib (SB1518), a JAK2/FLT3 inhibitor
for the treatment of acute myeloid leukemia.” Blood Cancer J 1 (11): e44. doi: 10.1038/bcj.2011.43.
Hatzimichael, E., E. Tsolas, and E. Briasoulis. 2014. “Profile of pacritinib and its potential in the
treatment of hematologic disorders.” J Blood Med 5:143–152. doi: 10.2147/JBM.S51253.
Hong, D., R. Kurzrock, Y. Kim, R. Woessner, A. Younes, J. Nemunaitis, N. Fowler, T. Zhou, J. Schmidt,
M. Jo, S. J. Lee, M. Yamashita, S. G. Hughes, L. Fayad, S. Piha-Paul, M. V. Nadella, M. Mohseni,
D. Lawson, C. Reimer, D. C. Blakey, X. Xiao, J. Hsu, A. Revenko, B. P. Monia, and
A. R. MacLeod. 2015. “AZD9150, a next-generation antisense oligonucleotide inhibitor of
STAT3 with early evidence of clinical activity in lymphoma and lung cancer.” Sci Transl Med 7
(314):314ra185. doi: 10.1126/scitranslmed.aac5272.
Hu, J., and W. X. Hu. 2018. “Targeting signaling pathways in multiple myeloma: pathogenesis and
implication for treatments.” Cancer Lett 414:214–221. doi: 10.1016/j.canlet.2017.11.020.
Huang, X., B. Meng, J. Iqbal, B. B. Ding, A. M. Perry, W. Cao, L. M. Smith, C. Bi, C. Jiang, T. C. Greiner,
D. D. Weisenburger, L. Rimsza, A. Rosenwald, G. Ott, J. Delabie, E. Campo, R. M. Braziel,
R. D. Gascoyne, J. R. Cook, R. R. Tubbs, E. S. Jaffe, J. O. Armitage, J. M. Vose, L. M. Staudt,
T. W. McKeithan, W. C. Chan, B. H. Ye, and K. Fu. 2013. “Activation of the STAT3 signaling
pathway is associated with poor survival in diffuse large B-cell lymphoma treated with R-CHOP.”
J Clin Oncol 31 (36):4520–4528. doi: 10.1200/JCO.2012.45.6004.
Ikezoe, T., S. Kojima, M. Furihata, J. Yang, C. Nishioka, A. Takeuchi, M. Isaka, H. P. Koeffler, and
A. Yokoyama. 2011. “Expression of p-JAK2 predicts clinical outcome and is a potential molecular
target of acute myelogenous leukemia.” Int J Cancer 129 (10):2512–2521. doi: 10.1002/ijc.25910.
James, C., V. Ugo, J. P. Le Couedic, J. Staerk, F. Delhommeau, C. Lacout, L. Garcon, H. Raslova,
R. Berger, A. Bennaceur-Griscelli, J. L. Villeval, S. N. Constantinescu, N. Casadevall, and
W. Vainchenker. 2005. “A unique clonal JAK2 mutation leading to constitutive signalling causes
polycythaemia vera.” Nature 434 (7037):1144–1148. doi: 10.1038/nature03546.
Joos, S., M. Kupper, S. Ohl, F. von Bonin, G. Mechtersheimer, M. Bentz, P. Marynen, P. Moller,
M. Pfreundschuh, L. Trumper, and P. Lichter. 2000. “Genomic imbalances including amplification
of the tyrosine kinase gene JAK2 in CD30+ Hodgkin cells.” Cancer Res 60 (3):549–552.
Jove, R. 2000. “Preface: STAT signaling.” Oncogene 19 (21):2466–2467. doi: 10.1038/sj.onc.1203549.
Kearney, L., D. Gonzalez De Castro, J. Yeung, J. Procter, S. W. Horsley, M. Eguchi-Ishimae,
C. M. Bateman, K. Anderson, T. Chaplin, B. D. Young, C. J. Harrison, H. Kempski, C. W. So,
A. M. Ford, and M. Greaves. 2009. “Specific JAK2 mutation (JAK2R683) and multiple gene
deletions in Down syndrome acute lymphoblastic leukemia.” Blood 113 (3):646–648. doi: 10.1182/
blood-2008-08-170928.
Klampfl, T., H. Gisslinger, A. S. Harutyunyan, H. Nivarthi, E. Rumi, J. D. Milosevic, N. C. Them, T. Berg,
B. Gisslinger, D. Pietra, D. Chen, G. I. Vladimer, K. Bagienski, C. Milanesi, I. C. Casetti,
E. Sant’Antonio, V. Ferretti, C. Elena, F. Schischlik, C. Cleary, M. Six, M. Schalling, A. Schonegger,
C. Bock, L. Malcovati, C. Pascutto, G. Superti-Furga, M. Cazzola, and R. Kralovics. 2013. “Somatic
mutations of calreticulin in myeloproliferative neoplasms.” N Engl J Med 369 (25):2379–2390. doi:
10.1056/NEJMoa1311347.
Kralovics, R., F. Passamonti, A. S. Buser, S. S. Teo, R. Tiedt, J. R. Passweg, A. Tichelli, M. Cazzola, and
R. C. Skoda. 2005. “A gain-of-function mutation of JAK2 in myeloproliferative disorders.” N Engl
J Med 352 (17):1779–1790. doi: 10.1056/NEJMoa051113.
Kroemer, G., L. Galluzzi, and L. Zitvogel. 2016. “STAT3 inhibition for cancer therapy: cell-autonomous
effects only?.” Oncoimmunology 5 (5):e1126063. doi: 10.1080/2162402X.2015.1126063.
Kuter, D. J. 2013. “The biology of thrombopoietin and thrombopoietin receptor agonists.” Int J Hematol
98 (1):10–23. doi: 10.1007/s12185-013-1382-0.
Kvasnicka, H. M., J. Thiele, C. E. Bueso-Ramos, W. Sun, J. Cortes, H. M. Kantarjian, and S. Verstovsek.
2018. “Long-term effects of ruxolitinib versus best available therapy on bone marrow fibrosis in
patients with myelofibrosis.” J Hematol Oncol 11 (1):42. doi: 10.1186/s13045-018-0585-5.
Lacronique, V., A. Boureux, V. D. Valle, H. Poirel, C. T. Quang, M. Mauchauffe, C. Berthou, M. Lessard,
R. Berger, J. Ghysdael, and O. A. Bernard. 1997. “A TEL-JAK2 fusion protein with constitutive kinase
activity in human leukemia.” Science 278 (5341):1309–1312. doi: 10.1126/science.278.5341.1309.
236 JAK-STAT Signaling in Diseases

Lee, D. S., R. A. O’Keefe, P. K. Ha, J. R. Grandis, and D. E. Johnson. 2018a. “Biochemical Properties of
a Decoy Oligodeoxynucleotide Inhibitor of STAT3 Transcription Factor.” Int J Mol Sci 19:6. doi:
10.3390/ijms19061608.
Lee, S., T. Shah, C. Yin, J. Hochberg, J. Ayello, E. Morris, C. van de Ven, and M. S. Cairo.
2018b. “Ruxolitinib significantly enhances in vitro apoptosis in Hodgkin lymphoma and primary
mediastinal B-cell lymphoma and survival in a lymphoma xenograft murine model.” Oncotarget 9
(11):9776–9788. doi: 10.18632/oncotarget.24267.
Leonard, W. J., and J. J. O’Shea. 1998. “Jaks and STATs: biological implications.” Annu Rev Immunol
16:293–322. doi: 10.1146/annurev.immunol.16.1.293.
Levine, R. L., M. Loriaux, B. J. Huntly, M. L. Loh, M. Beran, E. Stoffregen, R. Berger, J. J. Clark,
S. G. Willis, K. T. Nguyen, N. J. Flores, E. Estey, N. Gattermann, S. Armstrong, A. T. Look,
J. D. Griffin, O. A. Bernard, M. C. Heinrich, D. G. Gilliland, B. Druker, and M. W. Deininger.
2005a. “The JAK2V617F activating mutation occurs in chronic myelomonocytic leukemia and
acute myeloid leukemia, but not in acute lymphoblastic leukemia or chronic lymphocytic
leukemia.” Blood 106 (10):3377–3379. doi: 10.1182/blood-2005-05-1898.
Levine, R. L., M. Wadleigh, J. Cools, B. L. Ebert, G. Wernig, B. J. Huntly, T. J. Boggon, I. Wlodarska,
J. J. Clark, S. Moore, J. Adelsperger, S. Koo, J. C. Lee, S. Gabriel, T. Mercher, A. D’Andrea, S. Frohling,
K. Dohner, P. Marynen, P. Vandenberghe, R. A. Mesa, A. Tefferi, J. D. Griffin, M. J. Eck, W. R. Sellers,
M. Meyerson, T. R. Golub, S. J. Lee, and D. G. Gilliland. 2005b. “Activating mutation in the tyrosine
kinase JAK2 in polycythemia vera, essential thrombocythemia, and myeloid metaplasia with
myelofibrosis.” Cancer Cell 7 (4):387–397. doi: 10.1016/j.ccr.2005.03.023.
Levy, D. E., and J. E. Darnell, Jr. 2002. “Stats: transcriptional control and biological impact.” Nat Rev
Mol Cell Biol 3 (9):651–662. doi: 10.1038/nrm909.
Levy, D. E., and G. Inghirami. 2006. “STAT3: a multifaceted oncogene.” Proc Natl Acad Sci U S A 103
(27):10151–10152. doi: 10.1073/pnas.0604042103.
Liao, W., J. X. Lin, and W. J. Leonard. 2011. “IL-2 family cytokines: new insights into the complex roles
of IL-2 as a broad regulator of T helper cell differentiation.” Curr Opin Immunol 23 (5):598–604.
doi: 10.1016/j.coi.2011.08.003.
Liu, Y., and N. S. Gray. 2006. “Rational design of inhibitors that bind to inactive kinase conformations.”
Nat Chem Biol 2 (7):358–364. doi: 10.1038/nchembio799.
Lu, L., F. Zhu, M. Zhang, Y. Li, A. C. Drennan, S. Kimpara, I. Rumball, C. Selzer, H. Cameron,
A. Kellicut, A. Kelm, F. Wang, T. A. Waldmann, and L. Rui. 2018. “Gene regulation and
suppression of type I interferon signaling by STAT3 in diffuse large B cell lymphoma.” Proc Natl
Acad Sci U S A 115 (3):E498–E505. doi: 10.1073/pnas.1715118115.
Mandal, P. K., F. Gao, Z. Lu, Z. Ren, R. Ramesh, J. S. Birtwistle, K. K. Kaluarachchi, X. Chen,
R. C. Bast, Jr., W. S. Liao, and J. S. McMurray. 2011. “Potent and selective phosphopeptide mimetic
prodrugs targeted to the Src homology 2 (SH2) domain of signal transducer and activator of
transcription 3.” J Med Chem 54 (10):3549–3563. doi: 10.1021/jm2000882.
Mascarenhas, J., R. Hoffman, M. Talpaz, A. T. Gerds, B. Stein, V. Gupta, A. Szoke, M. Drummond,
A. Pristupa, T. Granston, R. Daly, S. Al-Fayoumi, J. A. Callahan, J. W. Singer, J. Gotlib,
C. Jamieson, C. Harrison, R. Mesa, and S. Verstovsek. 2018. “Pacritinib vs best available therapy,
including ruxolitinib, in patients with myelofibrosis: a randomized clinical trial.” JAMA Oncol 4
(5):652–659. doi: 10.1001/jamaoncol.2017.5818.
Mesa, R. A., J. J. Kiladjian, J. V. Catalano, T. Devos, M. Egyed, A. Hellmann, D. McLornan,
K. Shimoda, E. F. Winton, W. Deng, R. L. Dubowy, J. D. Maltzman, F. Cervantes, and J. Gotlib.
2017a. “SIMPLIFY-1: a phase iii randomized trial of momelotinib versus ruxolitinib in janus
kinase inhibitor-naive patients with myelofibrosis.” J Clin Oncol 35 (34):3844–3850. doi: 10.1200/
JCO.2017.73.4418.
Mesa, R. A., A. M. Vannucchi, A. Mead, M. Egyed, A. Szoke, A. Suvorov, J. Jakucs, A. Perkins,
R. Prasad, J. Mayer, J. Demeter, P. Ganly, J. W. Singer, H. Zhou, J. P. Dean, P. A. Te Boekhorst,
J. Nangalia, J. J. Kiladjian, and C. N. Harrison. 2017b. “Pacritinib versus best available therapy for
the treatment of myelofibrosis irrespective of baseline cytopenias (PERSIST-1): an international,
randomised, phase 3 trial.” Lancet Haematol 4 (5):e225–e236. doi: 10.1016/S2352-3026(17)30027-3.
Meyer, S. C., M. D. Keller, S. Chiu, P. Koppikar, O. A. Guryanova, F. Rapaport, K. Xu, K. Manova,
D. Pankov, R. J. O’Reilly, M. Kleppe, A. S. McKenney, A. H. Shih, K. Shank, J. Ahn, E. Papalexi,
JAK-STAT in Hematopoietic Malignancies 237

B. Spitzer, N. Socci, A. Viale, E. Mandon, N. Ebel, R. Andraos, J. Rubert, E. Dammassa,


V. Romanet, A. Dolemeyer, M. Zender, M. Heinlein, R. Rampal, R. S. Weinberg, R. Hoffman,
W. R. Sellers, F. Hofmann, M. Murakami, F. Baffert, C. Gaul, T. Radimerski, and R. L. Levine.
2015. “CHZ868, a Type II JAK2 inhibitor, reverses Type I JAK inhibitor persistence and
demonstrates efficacy in myeloproliferative neoplasms.” Cancer Cell 28 (1):15–28. doi: 10.1016/j.
ccell.2015.06.006.
Mitsiades, C. S., N. S. Mitsiades, N. C. Munshi, P. G. Richardson, and K. C. Anderson. 2006. “The role of
the bone microenvironment in the pathophysiology and therapeutic management of multiple
myeloma: interplay of growth factors, their receptors and stromal interactions.” Eur J Cancer 42
(11):1564–1573. doi: 10.1016/j.ejca.2005.12.025.
Moulis, G., A. Palmaro, J. L. Montastruc, B. Godeau, M. Lapeyre-Mestre, and L. Sailler. 2014.
“Epidemiology of incident immune thrombocytopenia: a nationwide population-based study in
France.” Blood 124 (22):3308–3315. doi: 10.1182/blood-2014-05-578336.
Mullighan, C. G., J. Zhang, R. C. Harvey, J. R. Collins-Underwood, B. A. Schulman, L. A. Phillips,
S. K. Tasian, M. L. Loh, X. Su, W. Liu, M. Devidas, S. R. Atlas, I. M. Chen, R. J. Clifford,
D. S. Gerhard, W. L. Carroll, G. H. Reaman, M. Smith, J. R. Downing, S. P. Hunger, and
C. L. Willman. 2009. “JAK mutations in high-risk childhood acute lymphoblastic leukemia.” Proc
Natl Acad Sci U S A 106 (23):9414–9418. doi: 10.1073/pnas.0811761106.
Murati, A., V. Gelsi-Boyer, J. Adelaide, C. P erot, P. Talmant, S. Giraudier, L. Lode, A. Letessier,
B. Delaval, V. Brunel, M. Imbert, R. Garand, L. Xerri, D. Birnbaum, M. J. Mozziconacci, and
M. Chaffanet. 2005. “PCM1-JAK2 fusion in myeloproliferative disorders and acute erythroid
leukemia with t(8;9) translocation.” Leukemia 19 (9):1692–1696. doi: 10.1038/sj.leu.2403879.
Nangalia, J., C. E. Massie, E. J. Baxter, F. L. Nice, G. Gundem, D. C. Wedge, E. Avezov, J. Li, K. Kollmann,
D. G. Kent, A. Aziz, A. L. Godfrey, J. Hinton, I. Martincorena, P. Van Loo, A. V. Jones, P. Guglielmelli,
P. Tarpey, H. P. Harding, J. D. Fitzpatrick, C. T. Goudie, C. A. Ortmann, S. J. Loughran, K. Raine,
D. R. Jones, A. P. Butler, J. W. Teague, S. O’Meara, S. McLaren, M. Bianchi, Y. Silber,
D. Dimitropoulou, D. Bloxham, L. Mudie, M. Maddison, B. Robinson, C. Keohane, C. Maclean,
K. Hill, K. Orchard, S. Tauro, M. Q. Du, M. Greaves, D. Bowen, B. J. P. Huntly, C. N. Harrison,
N. C. P. Cross, D. Ron, A. M. Vannucchi, E. Papaemmanuil, P. J. Campbell, and A. R. Green. 2013.
“Somatic CALR mutations in myeloproliferative neoplasms with nonmutated JAK2.” N Engl J Med
369 (25):2391–2405. doi: 10.1056/NEJMoa1312542.
Navarro, A., T. Diaz, A. Martinez, A. Gaya, A. Pons, B. Gel, C. Codony, G. Ferrer, C. Martinez,
E. Montserrat, and M. Monzo. 2009. “Regulation of JAK2 by miR-135a: prognostic impact in
classic Hodgkin lymphoma.” Blood 114 (14):2945–2951. doi: 10.1182/blood-2009-02-204842.
Nebral, K., D. Denk, A. Attarbaschi, M. Konig, G. Mann, O. A. Haas, and S. Strehl. 2009. “Incidence
and diversity of PAX5 fusion genes in childhood acute lymphoblastic leukemia.” Leukemia 23
(1):134–143. doi: 10.1038/leu.2008.306.
Nelson, E. A., S. R. Walker, A. Kepich, L. B. Gashin, T. Hideshima, H. Ikeda, D. Chauhan,
K. C. Anderson, and D. A. Frank. 2008. “Nifuroxazide inhibits survival of multiple myeloma cells
by directly inhibiting STAT3.” Blood 112 (13):5095–5102. doi: 10.1182/blood-2007-12-129718.
Nelson, E. A., S. R. Walker, E. Weisberg, M. Bar-Natan, R. Barrett, L. B. Gashin, S. Terrell,
J. L. Klitgaard, L. Santo, M. R. Addorio, B. L. Ebert, J. D. Griffin, and D. A. Frank. 2011. “The
STAT5 inhibitor pimozide decreases survival of chronic myelogenous leukemia cells resistant to
kinase inhibitors.” Blood 117 (12):3421–3429. doi: 10.1182/blood-2009-11-255232.
Nelson, E. A., S. R. Walker, M. Xiang, E. Weisberg, M. Bar-Natan, R. Barrett, S. Liu, S. Kharbanda,
A. L. Christie, M. Nicolais, J. D. Griffin, R. M. Stone, A. L. Kung, and D. A. Frank. 2012. “The
STAT5 inhibitor pimozide displays efficacy in models of acute myelogenous leukemia driven by
FLT3 mutations.” Genes Cancer 3 (7–8):503–511. doi: 10.1177/1947601912466555.
Nielsen, C., S. E. Bojesen, B. G. Nordestgaard, K. F. Kofoed, and H. S. Birgens. 2014. “JAK2V617F
somatic mutation in the general population: myeloproliferative neoplasm development and pro­
gression rate.” Haematologica 99 (9):1448–1455. doi: 10.3324/haematol.2014.107631.
Ok, C. Y., J. Chen, Z. Y. Xu-Monette, A. Tzankov, G. C. Manyam, L. Li, C. Visco, S. Montes-Moreno,
K. Dybkaer, A. Chiu, A. Orazi, Y. Zu, G. Bhagat, K. L. Richards, E. D. Hsi, W. W. Choi, J. H. van
Krieken, J. Huh, X. Zhao, M. Ponzoni, A. J. Ferreri, F. Bertoni, J. P. Farnen, M. B. Moller,
M. A. Piris, J. N. Winter, L. J. Medeiros, and K. H. Young. 2014. “Clinical implications of
238 JAK-STAT Signaling in Diseases

phosphorylated STAT3 expression in De Novo diffuse large B-cell lymphoma.” Clin Cancer Res 20
(19):5113–5123. doi: 10.1158/1078-0432.CCR-14-0683.
O’Shea, J. J., D. M. Schwartz, A. V. Villarino, M. Gadina, I. B. McInnes, and A. Laurence. 2015. “The
JAK-STAT pathway: impact on human disease and therapeutic intervention.” Annu Rev Med
66:311–328. doi: 10.1146/annurev-med-051113-024537.
Padron, E., A. Dezern, M. Andrade-Campos, K. Vaddi, P. Scherle, Q. Zhang, Y. Ma, M. E. Balasis,
S. Tinsley, H. Ramadan, C. Zimmerman, D. P. Steensma, G. J. Roboz, J. E. Lancet, A. F. List,
M. A. Sekeres, R. S. Komrokji, and Consortium Myelodysplastic Syndrome Clinical Research.
2016. “A multi-institution phase i trial of ruxolitinib in patients with chronic myelomonocytic
Leukemia (CMML).” Clin Cancer Res 22 (15):3746–3754. doi: 10.1158/1078-0432.CCR-15-2781.
Padron, E., J. S. Painter, S. Kunigal, A. W. Mailloux, K. McGraw, J. M. McDaniel, E. Kim,
C. Bebbington, M. Baer, G. Yarranton, J. Lancet, R. S. Komrokji, O. Abdel-Wahab, A. F. List,
and P. K. Epling-Burnette. 2013. “GM-CSF-dependent pSTAT5 sensitivity is a feature with
therapeutic potential in chronic myelomonocytic leukemia.” Blood 121 (25):5068–5077. doi:
10.1182/blood-2012-10-460170.
Pan, Y. R., C. C. Chen, Y. T. Chan, H. J. Wang, F. T. Chien, Y. L. Chen, J. L. Liu, and M. H. Yang. 2018.
“STAT3-coordinated migration facilitates the dissemination of diffuse large B-cell lymphomas.”
Nat Commun 9 (1):3696. doi: 10.1038/s41467-018-06134-z.
Pardanani, A., R. R. Laborde, T. L. Lasho, C. Finke, K. Begna, A. Al-Kali, W. J. Hogan, M. R. Litzow,
A. Leontovich, M. Kowalski, and A. Tefferi. 2013. “Safety and efficacy of CYT387, a JAK1 and
JAK2 inhibitor, in myelofibrosis.” Leukemia 27 (6):1322–1327. doi: 10.1038/leu.2013.71.
Pardanani, A., T. Lasho, G. Smith, C. J. Burns, E. Fantino, and A. Tefferi. 2009. “CYT387, a selective JAK1/
JAK2 inhibitor: in vitro assessment of kinase selectivity and preclinical studies using cell lines and
primary cells from polycythemia vera patients.” Leukemia 23 (8):1441–1445. doi: 10.1038/leu.2009.50.
Pardanani, A. D., R. L. Levine, T. Lasho, Y. Pikman, R. A. Mesa, M. Wadleigh, D. P. Steensma,
M. A. Elliott, A. P. Wolanskyj, W. J. Hogan, R. F. McClure, M. R. Litzow, D. G. Gilliland, and
A. Tefferi. 2006. “MPL515 mutations in myeloproliferative and other myeloid disorders: a study of
1182 patients.” Blood 108 (10):3472–3476. doi: 10.1182/blood-2006-04-018879.
Passamonti, F., M. Griesshammer, F. Palandri, M. Egyed, G. Benevolo, T. Devos, J. Callum,
A. M. Vannucchi, S. Sivgin, C. Bensasson, M. Khan, N. Mounedji, and G. Saydam. 2017.
“Ruxolitinib for the treatment of inadequately controlled polycythaemia vera without splenomegaly
(RESPONSE-2): a randomised, open-label, phase 3b study.” Lancet Oncol 18 (1):88–99. doi: 10.1016/
S1470-2045(16)30558-7.
Patnaik, M. M., R. A. Knudson, N. Gangat, C. A. Hanson, A. Pardanani, A. Tefferi, and
R. P. Ketterling. 2010. “Chromosome 9p24 abnormalities: prevalence, description of novel JAK2
translocations, JAK2V617F mutation analysis and clinicopathologic correlates.” Eur J Haematol
84 (6):518–524. doi: 10.1111/j.1600-0609.2010.01428.x.
Pianko, M. J., Y. Liu, S. Bagchi, and A. M. Lesokhin. 2017. “Immune checkpoint blockade for
hematologic malignancies: a review.” Stem Cell Investig 4:32. doi: 10.21037/sci.2017.03.04.
Podar, K., D. Chauhan, and K. C. Anderson. 2009. “Bone marrow microenvironment and the identification
of new targets for myeloma therapy.” Leukemia 23 (1):10–24. doi: 10.1038/leu.2008.259.
Poitras, J. L., P. Dal Cin, J. C. Aster, D. J. Deangelo, and C. C. Morton. 2008. “Novel SSBP2-JAK2 fusion
gene resulting from a t(5;9)(q14.1;p24.1) in pre-B acute lymphocytic leukemia.” Genes Chromo­
somes Cancer 47 (10):884–889. doi: 10.1002/gcc.20585.
Poulsen, A., A. William, S. Blanchard, A. Lee, H. Nagaraj, H. Wang, E. Teo, E. Tan, K. C. Goh, and
B. Dymock. 2012. “Structure-based design of oxygen-linked macrocyclic kinase inhibitors: discov­
ery of SB1518 and SB1578, potent inhibitors of Janus kinase 2 (JAK2) and Fms-like tyrosine
kinase-3 (FLT3).” J Comput Aided Mol Des 26 (4):437–450. doi: 10.1007/s10822-012-9572-z.
Rampal, R., F. Al-Shahrour, O. Abdel-Wahab, J. P. Patel, J. P. Brunel, C. H. Mermel, A. J. Bass, J. Pretz,
J. Ahn, T. Hricik, O. Kilpivaara, M. Wadleigh, L. Busque, D. G. Gilliland, T. R. Golub, B. L. Ebert,
and R. L. Levine. 2014. “Integrated genomic analysis illustrates the central role of JAK-STAT
pathway activation in myeloproliferative neoplasm pathogenesis.” Blood 123 (22):e123–e133. doi:
10.1182/blood-2014-02-554634.
Reiter, A., C. Walz, A. Watmore, C. Schoch, I. Blau, B. Schlegelberger, U. Berger, N. Telford, S. Aruliah,
J. A. Yin, D. Vanstraelen, H. F. Barker, P. C. Taylor, A. O’Driscoll, F. Benedetti, C. Rudolph,
JAK-STAT in Hematopoietic Malignancies 239

H. J. Kolb, A. Hochhaus, R. Hehlmann, A. Chase, and N. C. Cross. 2005. “The t(8;9)(p22;p24) is


a recurrent abnormality in chronic and acute leukemia that fuses PCM1 to JAK2.” Cancer Res 65
(7):2662–2667. doi: 10.1158/0008-5472.CAN-04-4263.
Rosenwald, A., G. Wright, K. Leroy, X. Yu, P. Gaulard, R. D. Gascoyne, W. C. Chan, T. Zhao,
C. Haioun, T. C. Greiner, D. D. Weisenburger, J. C. Lynch, J. Vose, J. O. Armitage, E. B. Smeland,
S. Kvaloy, H. Holte, J. Delabie, E. Campo, E. Montserrat, A. Lopez-Guillermo, G. Ott,
H. K. Muller-Hermelink, J. M. Connors, R. Braziel, T. M. Grogan, R. I. Fisher, T. P. Miller,
M. LeBlanc, M. Chiorazzi, H. Zhao, L. Yang, J. Powell, W. H. Wilson, E. S. Jaffe, R. Simon,
R. D. Klausner, and L. M. Staudt. 2003. “Molecular diagnosis of primary mediastinal B cell
lymphoma identifies a clinically favorable subgroup of diffuse large B cell lymphoma related to
Hodgkin lymphoma.” J Exp Med 198 (6):851–862. doi: 10.1084/jem.20031074.
Saharinen, P., and O. Silvennoinen. 2002. “The pseudokinase domain is required for suppression of basal
activity of Jak2 and Jak3 tyrosine kinases and for cytokine-inducible activation of signal
transduction.” J Biol Chem 277 (49):47954–47963. doi: 10.1074/jbc.M205156200.
Saharinen, P., M. Vihinen, and O. Silvennoinen. 2003. “Autoinhibition of Jak2 tyrosine kinase is
dependent on specific regions in its pseudokinase domain.” Mol Biol Cell 14 (4):1448–1459. doi:
10.1091/mbc.e02-06-0342.
Sansone, P., and J. Bromberg. 2012. “Targeting the interleukin-6/Jak/stat pathway in human
malignancies.” J Clin Oncol 30 (9):1005–1014. doi: 10.1200/JCO.2010.31.8907.
Scott, L. M., P. J. Campbell, E. J. Baxter, T. Todd, P. Stephens, S. Edkins, R. Wooster, M. R. Stratton,
P. A. Futreal, and A. R. Green. 2005. “The V617F JAK2 mutation is uncommon in cancers and in
myeloid malignancies other than the classic myeloproliferative disorders.” Blood 106 (8):2920–2921.
doi: 10.1182/blood-2005-05-2087.
Seif, F., M. Khoshmirsafa, H. Aazami, M. Mohsenzadegan, G. Sedighi, and M. Bahar. 2017. “The role of
JAK-STAT signaling pathway and its regulators in the fate of T helper cells.” Cell Commun Signal
15 (1):23. doi: 10.1186/s12964-017-0177-y.
Seth, P. P., G. Vasquez, C. A. Allerson, A. Berdeja, H. Gaus, G. A. Kinberger, T. P. Prakash,
M. T. Migawa, B. Bhat, and E. E. Swayze. 2010. “Synthesis and biophysical evaluation of 2ʹ,4ʹ­
constrained 2ʹO-methoxyethyl and 2ʹ,4ʹ-constrained 2ʹO-ethyl nucleic acid analogues.” J Org Chem
75 (5):1569–1581. doi: 10.1021/jo902560f.
Silvennoinen, O., and S. R. Hubbard. 2015. “Targeting the inactive conformation of JAK2 in hematolo­
gical malignancies.” Cancer Cell 28 (1):1–2. doi: 10.1016/j.ccell.2015.06.010.
Skoda, R. C., A. Duek, and J. Grisouard. 2015. “Pathogenesis of myeloproliferative neoplasms.” Exp
Hematol 43 (8):599–608. doi: 10.1016/j.exphem.2015.06.007.
Solary, E. 2016. “Unplugging JAK/STAT in chronic myelomonocytic Leukemia.” Clin Cancer Res 22
(15):3707–3709. doi: 10.1158/1078-0432.CCR-16-0372.
Steensma, D. P., R. F. McClure, J. E. Karp, A. Tefferi, T. L. Lasho, H. L. Powell, G. W. DeWald, and
S. H. Kaufmann. 2006. “JAK2 V617F is a rare finding in de novo acute myeloid leukemia, but
STAT3 activation is common and remains unexplained.” Leukemia 20 (6):971–978. doi: 10.1038/sj.
leu.2404206.
Syed, Y. Y. 2017. “Durvalumab: first global approval.” Drugs 77 (12):1369–1376. doi: 10.1007/s40265­
017-0782-5.
Tyner, J. W., T. G. Bumm, J. Deininger, L. Wood, K. J. Aichberger, M. M. Loriaux, B. J. Druker,
C. J. Burns, E. Fantino, and M. W. Deininger. 2010. “CYT387, a novel JAK2 inhibitor, induces
hematologic responses and normalizes inflammatory cytokines in murine myeloproliferative
neoplasms.” Blood 115 (25):5232–5240. doi: 10.1182/blood-2009-05-223727.
Vainchenker, W., and S. N. Constantinescu. 2013. “JAK/STAT signaling in hematological malignancies.”
Oncogene 32 (21):2601–2613. doi: 10.1038/onc.2012.347.
Van Den Neste, E., M. Andre, T. Gastinne, A. Stamatoullas, C. Haioun, A. Belhabri, O. Reman,
O. Casasnovas, H. Ghesquieres, G. Verhoef, M. J. Claessen, H. A. Poirel, M. C. Copin, R. Dubois,
P. Vandenberghe, I. A. Stoian, A. S. Cottereau, S. Bailly, L. Knoops, and F. Morschhauser. 2018.
“A phase II study of the oral JAK1/JAK2 inhibitor ruxolitinib in advanced relapsed/refractory
Hodgkin lymphoma.” Haematologica 103 (5):840–848. doi: 10.3324/haematol.2017.180554.
Vannucchi, A. M. 2015. “Ruxolitinib versus standard therapy for the treatment of polycythemia vera.”
N Engl J Med 372 (17):1670–1671. doi: 10.1056/NEJMc1502524.
240 JAK-STAT Signaling in Diseases

Vannucchi, A. M., S. Verstovsek, P. Guglielmelli, M. Griesshammer, T. C. Burn, A. Naim,


D. Paranagama, M. Marker, B. Gadbaw, and J. J. Kiladjian. 2017. “Ruxolitinib reduces JAK2 p.
V617F allele burden in patients with polycythemia vera enrolled in the RESPONSE study.” Ann
Hematol 96 (7):1113–1120. doi: 10.1007/s00277-017-2994-x.
Van Roosbroeck, K., L. Cox, T. Tousseyn, I. Lahortiga, O. Gielen, B. Cauwelier, P. De Paepe, G. Verhoef,
P. Marynen, P. Vandenberghe, C. De Wolf-Peeters, J. Cools, and I. Wlodarska. 2011. “JAK2
rearrangements, including the novel SEC31A-JAK2 fusion, are recurrent in classical Hodgkin
lymphoma.” Blood 117 (15):4056–4064. doi: 10.1182/blood-2010-06-291310.
Velazquez, L., M. Fellous, G. R. Stark, and S. Pellegrini. 1992. “A protein tyrosine kinase in the
interferon alpha/beta signaling pathway.” Cell 70 (2):313–322. doi: 10.1016/0092-8674(92)90105-l.
Verstovsek, S., J. Gotlib, R. A. Mesa, A. M. Vannucchi, J. J. Kiladjian, F. Cervantes, C. N. Harrison,
R. Paquette, W. Sun, A. Naim, P. Langmuir, T. Dong, P. Gopalakrishna, and V. Gupta. 2017.
“Long-term survival in patients treated with ruxolitinib for myelofibrosis: COMFORT-I and -II
pooled analyses.” J Hematol Oncol 10 (1):156. doi: 10.1186/s13045-017-0527-7.
Verstovsek, S., R. A. Mesa, J. Gotlib, R. S. Levy, V. Gupta, J. F. DiPersio, J. V. Catalano, M. Deininger,
C. Miller, R. T. Silver, M. Talpaz, E. F. Winton, J. H. Harvey, Jr., M. O. Arcasoy, E. Hexner,
R. M. Lyons, R. Paquette, A. Raza, K. Vaddi, S. Erickson-Viitanen, I. L. Koumenis, W. Sun,
V. Sandor, and H. M. Kantarjian. 2012. “A double-blind, placebo-controlled trial of ruxolitinib for
myelofibrosis.” N Engl J Med 366 (9):799–807. doi: 10.1056/NEJMoa1110557.
Verstovsek, S., A. M. Vannucchi, M. Griesshammer, T. Masszi, S. Durrant, F. Passamonti,
C. N. Harrison, F. Pane, P. Zachee, K. Kirito, C. Besses, M. Hino, B. Moiraghi, C. B. Miller,
M. Cazzola, V. Rosti, I. Blau, R. Mesa, M. M. Jones, H. Zhen, J. Li, N. Francillard, D. Habr, and
J. J. Kiladjian. 2016. “Ruxolitinib versus best available therapy in patients with polycythemia vera:
80-week follow-up from the RESPONSE trial.” Haematologica 101 (7):821–829. doi: 10.3324/
haematol.2016.143644.
Wardrop, D., and D. P. Steensma. 2009. “Is refractory anaemia with ring sideroblasts and thrombocytosis
(RARS-T) a necessary or useful diagnostic category?.” Br J Haematol 144 (6):809–817. doi:
10.1111/j.1365-2141.2008.07526.x.
Weisberg, E., Q. Liu, E. Nelson, A. L. Kung, A. L. Christie, R. Bronson, M. Sattler, T. Sanda, Z. Zhao,
W. Hur, C. Mitsiades, R. Smith, J. F. Daley, R. Stone, I. Galinsky, J. D. Griffin, and N. Gray. 2012.
“Using combination therapy to override stromal-mediated chemoresistance in mutant
FLT3-positive AML: synergism between FLT3 inhibitors, dasatinib/multi-targeted inhibitors and
JAK inhibitors.” Leukemia 26 (10):2233–2244. doi: 10.1038/leu.2012.96.
Xu, X., Q. Zhang, J. Luo, S. Xing, Q. Li, S. B. Krantz, X. Fu, and Z. J. Zhao. 2007. “JAK2(V617F):
prevalence in a large Chinese hospital population.” Blood 109 (1):339–342. doi: 10.1182/blood­
2006-03-009472.
Younes, A., J. Romaguera, M. Fanale, P. McLaughlin, F. Hagemeister, A. Copeland, S. Neelapu,
L. Kwak, J. Shah, S. de Castro Faria, S. Hart, J. Wood, R. Jayaraman, K. Ethirajulu, and J. Zhu.
2012. “Phase I study of a novel oral Janus kinase 2 inhibitor, SB1518, in patients with relapsed
lymphoma: evidence of clinical and biologic activity in multiple lymphoma subtypes.” J Clin Oncol
30 (33):4161–4167. doi: 10.1200/JCO.2012.42.5223.
Yu, H., D. Pardoll, and R. Jove. 2009. “STATs in cancer inflammation and immunity: a leading role for
STAT3.” Nat Rev Cancer 9 (11):798–809. doi: 10.1038/nrc2734.
Zhang, Q., H. Bai, X. H. Yu, B. Wu, Y. Z. Pan, C. B. Wang, L. P. Zhao, W. B. Li, F. Xu, and J. Zhang.
2018. “[Changes of IL-21 and its mediated JAK/STAT signaling pathway in patients with immune
thrombocytopenia].” Zhongguo Shi Yan Xue Ye Xue Za Zhi 26 (3):859–865. doi: 10.7534/j.
issn.1009-2137.2018.03.038.
Zhang, Y., and J. M. Kolesar. 2011. “Eltrombopag: an oral thrombopoietin receptor agonist for the
treatment of idiopathic thrombocytopenic purpura.” Clin Ther 33 (11):1560–1576. doi: 10.1016/j.
clinthera.2011.10.004.
13
Immunodeficiency: Consequences of Mutations in
JAK-STAT Signaling

Daniel Silberger and Duy Pham


Department of Pathology
University of Alabama at Birmingham
Birmingham, Alabama

13.1 Overview JAK-STAT Signaling


The JAK-STAT is a classical pathway that plays an essential role in immunity, cell division, cell death,
and tumor formation (Stark and Darnell 2012). JAK-STAT signaling starts with the interaction
between extracellular stimuli such as cytokines or growth factors with their corresponding transmem­
brane receptors (Murray 2007). Cytokine-receptor interaction induces receptor dimerization, which
promotes transactivation of receptor-associated JAKs by bringing it into close proximity and by
prompting conformational changes that separate their kinase domains and inhibitory pseudo-kinase
domains (Leonard 2002; Rawlings 2004; Schindler 1999). Activated JAKs then phosphorylate the
receptor’s tyrosine residues to create docking sites for SH2 domain-associated STAT proteins (Meyer
et al. 2004). Subsequently, JAKs phosphorylate receptor-bound STATs, resulting in STAT dissocia­
tion from the receptor. Activated STATs can form homo- or heterodimers with other STATs,
translocate to the cell nucleus, bind DNA, and promote transcription of target genes (Figure 13.1).

FIGURE 13.1 JAK-STAT signaling cascade. (1) Cytokine-receptor interaction and receptor dimerization (2) transac­
tivation of receptor-associated JAKs (3) JAK phosphorylation of receptor (4) STAT binds phosphorylated receptor
and is phosphorylated by JAK (5) STAT dimerization, and (6) STAT dimers translocate to nucleus, bind DNA, and
modulate gene expression

241
242 JAK-STAT Signaling in Diseases

In mammals, there are four JAKs (JAK1, JAK2, JAK3, and TYK2) and seven STATs (STAT1,
STAT2, STAT3, STAT4, STAT5a, STAT5b, and STAT6) that are associated with the signaling of more
than 50 cytokines and growth factors (Bousoik and Aliabadi 2018; Darnell 1997). Importantly, the
specificity of JAK-STAT signaling greatly depends on cell lineages, qualitative differences in the
duration and or intensity, heterogeneous signaling involving in multiple STATs, and hierarchical order
of STATs in response to stimuli (O’Shea and Murray 2008; Seidel et al. 1995; Villarino et al. 2003).
As mentioned, STATs not only form homodimers and heterodimers in a stimuli-dependent
manner, but also form higher order tetramers that are poorly understood (Lin et al. 2017, 2012).
In contrast to dimerization that occurs in the cytoplasm, tetramerization assembles in the nucleus
upon binding of STAT dimers and DNA (John et al. 1999). For example, while Type I IFN
promotes STAT1/STAT2 heterodimers, Type II IFN elicits STAT1 homodimers (Au-Yeung et al.
2014). IL-10 also induces heterodimers between STAT1 and STAT3 at the sis oncogene (SIE)
sequence of the c-fos promoter (Wehinger et al. 1999). Another non-canonical model of JAK-STAT
signaling is JAK-independent activation of STAT (Nan, Wu, and Zhang 2018; Singh 2011). It has
been shown that certain receptor tyrosine kinases can bypass JAK activity and mediate the
phosphorylation of STAT on its own (Zhang et al. 2000). Flt-3 receptor and STING have been
demonstrated to activate STAT5 and STAT6, respectively, in a JAK-independent manner (Zhang
et al. 2000). In addition to tyrosine phosphorylation, other post-translational modifications regulate
STAT function, including serine phosphorylation, which influences the DNA-binding affinity of
STATs and the interaction of STATs with other factors (Lim and Cao 2006). STAT function can
also be altered by other modifications such as acetylation, methylation, and sumoylation, that can
either promote or limit STAT function (Mowen et al. 2001; Ungureanu et al. 2003; Yuan 2005). It
has been shown that both lysine acetylation and arginine methylation of STAT proteins can
enhance its function (Mowen et al. 2001; Zhuang 2013). In contrast, sumoylation has been strongly
associated with repressing the activity of STAT, including STAT1 and STAT5 (Begitt et al. 2011;
Van Nguyen et al. 2012). Another non-canonical characteristic of JAK-STAT is the location of its
function. Although STATs have been shown to localize to the nucleus and regulate gene expression,
recent studies have demonstrated its function elsewhere (Meier and Larner 2014). For example, all
other STATs with the exception of STAT4 have been reported to translocate to the mitochondrion,
but their function within this compartment remains poorly understood (Meier and Larner 2014).
STAT proteins act as classical transcription factors (TFs) with the ability to recognize and bind to
consensus DNA sequence known as GAS elements (Villarino et al. 2015). In T helper cells, STAT
binding has been mapped to various regions of the genome including proximal to transcription start
sites, enhancers, and other cis regulatory elements (Vahedi et al. 2012). However, each STAT has some
selectivity in the core sequence that results in variable DNA-binding affinity and overlapping binding
sites between individual STATs (O’Shea et al. 2011). This partially explains for common genes that are
regulated by multiple STAT family members (Villarino et al. 2015). In addition, STAT proteins are
capable of recruiting co-activators and other TFs to promote epigenetic changes and enhance transcrip­
tion of target genes (Shuai 2000; Vahedi et al. 2012). STATs have been shown to associate with CBP,
p300, and histone acetyltransferases (HATs) that promote active histone marks thus enhancing gene
expression (Shuai 2000). In addition, STATs integrate with other TFs to create regulatory networks that
have been dissected in detail in T cell biology (O’Shea et al. 2011; Vahedi et al. 2012). For example,
STAT3 involves in a transcriptional network including Irf4, Batf, and Rorγt that plays a crucial role in
T cell biology (Ciofani et al. 2012). Other groups show the cooperation of STATs with lineage-specific
TFs including T-bet, Gata-3, and Foxp3 (Maier et al. 2012; Passerini et al. 2008; Thieu et al. 2008).
Whether STATs have a central role in the network or are just components remain to be explored.

13.2 Regulation of JAK-STAT Signaling


Given the importance of the JAK-STAT pathway, every step of the signaling is strictly monitored.
One key component of the JAK-STAT signaling pathway is the phosphorylation of tyrosine
Immunodeficiency 243

residues of JAKs, STATs, and upstream receptors. Protein tyrosine phosphatases (PTPs), such as
SHP1, SHP2, and CD45, can hydrolyze phosphorylated tyrosine residues thus inhibiting JAK-STAT
signaling (Böhmer and Friedrich 2014; Shuai and Liu 2003). There are other PTPs including PTPRD,
PTPRT, TCPTP, PTP1B, and DUSP2 (Lu et al. 2015). Another key regulatory module is protein
inhibitors of activated STATs (PIAS) including four-member protein family: PIAS1, PIAS3, PIASx,
and PIASγ (Niu et al. 2018). These PIAS exert a variety of mechanisms to suppress STAT activation,
such as adding small ubiquitin-like modifier (SUMO) onto JAK and STAT to modify their function,
directly interacting with STATs to block their DNA-binding activity, and recruiting histone deacety­
lases (HDACs) that modulate gene expression (Niu et al. 2018). Another essential regulatory
component comprises proteins of the suppressor of cytokine signaling (SOCS) family including
cytokine-inducible SH2 domain-containing protein (CISH), SOCS1, SOCS2, SOCS3, SOCS4,
SOCS5, SOCS6, and SOCS7, but only the first four members are linked with the JAK-STAT pathway
in mammals (Croker et al. 2008; Palmer and Restifo 2009). SOCS proteins have a SH2 domain that
allows receptor binding and directly suppresses JAK activation (Croker et al. 2008). Since STATs can
induce the expression of SOCS, they act as negative feedback loops to modulate both quantitative and
qualitative aspects of JAK-STAT signaling. Last, but not least, STATs are maintained at a limiting
resource via self-regulated gene expression, post-transcriptional regulation, and post-translational
regulation, such that minor changes can have significant effects in its function (Qi and Yang 2014).

13.3 JAK-STAT in Immunology


13.3.1 T Helper Cell Development
T helper cells secrete cytokines that are important in mediating immune responses, and JAK­
STAT signaling is an essential component for the development of T cell lineages (Figure 13.2).
Th1 cells require T-bet, IFN-γ-STAT1, and IL-12-STAT4 for its development, which
produce IFN-γ and mediate immune response against intracellular pathogens (Athie-Morales
et al. 2004; G Ler et al. 1996; Ma et al. 2010). Th2 cells produce IL-4, IL-5, and IL-13 that
are important in mounting immune responses against parasites and contributing to allergic
reactions (Walker and McKenzie 2017). IL-4/STAT6 is required for the differentiation of Th2
cells and inhibition of Th1 cell differentiation (Kaplan et al. 1996; Wurtz 2004). Th17 cells
produce IL-17A, IL-17F, IL-21, and IL-22 that are critical in immunity against extracellular
bacteria and antifungal responses (Becher and Segal 2011). T follicular helper cells (Tfh)
produce IL-4 and IL-21 and are important in the maintenance of germinal center responses
(Crotty 2011). Both Th17 and Tfh cells require IL-6-STAT3 signaling for their development
(Foley 2007; H. Wu et al. 2016). Regulatory T cells (Treg) are essential to suppress immune
response to maintain homeostasis and self-tolerance (Bluestone and Tang 2005). IL-2-STAT5
is required for the generation of Treg cells, by promoting the expression of Foxp3, a Treg­
defining factor (Mahmud et al. 2013). Th9 cells secrete IL-9 and are essential in protecting
against helminth infections, promoting allergic responses, autoimmunity, and tumor suppres­
sion (Kaplan et al. 2015). Several STAT proteins are required to modulate Th9 cell differ­
entiation including STAT5 and STAT6 that act as positive regulators, while STAT3 inhibits
Th9 differentiation (Goswami et al. 2012; Olson et al. 2016; Ulrich et al. 2017). STATs play
a complicated role in the regulation of T helper cell differentiation. In fact, multiple STAT signals are
integrated in order to regulate the development of T cells (O’Shea et al. 2011). IL-2-STAT5 is also
required for Th1 and Th2 cell differentiation by promoting T-bet expression and IL4Rα-induced
GATA3 expression, respectively (Liao et al. 2008). In addition, STAT3 has been shown to be required
for Th2 cell differentiation (Stritesky et al. 2011). Type I IFN activates STAT1 and STAT2 that can
suppress Treg cell development (Srivastava et al. 2014). In addition, Type I IFN can induce STAT4
activation with lesser potency compared to IL-12 and might regulate Th1 cell differentiation (Ramos
et al. 2007). Taken together, JAK-STAT signaling plays a key role in T helper cell differentiation but
the interplay between STAT proteins is still poorly understood.
244

FIGURE 13.2 The contribution of JAK-STAT signaling in the development of T helper cell and ILC subsets. Cytokines activate JAK-STAT pathways that are essential in the
differentiation of immune cells
JAK-STAT Signaling in Diseases
Immunodeficiency 245

13.3.2 Innate Lymphoid Cell Development


JAK-STAT signaling also plays an essential role in the development of innate lymphoid cells
(ILC) (Figure 13.2) (Stabile et al. 2018). ILC1 cells are required for protective immunity but are
also involved in the homeostatic regulation of immunity. IL-7 and IL-15 provide critical signals
that are necessary for the development of ILC1 cells (Villarino et al. 2017b). This is further
demonstrated by mice with deficiencies in IL-15 or IL-2Rγ, which have reduced ILC1 subsets
(Lodolce et al. 1998; Ranson et al. 2003). In addition, IL-7 signaling is pivotal for the develop­
ment of Type 2 (ILC2) and type 3 (ILC3) ILCs cells (Vonarbourg and Diefenbach 2012). Mice
with deficiency in IL-7 signaling, including Il7ra−/− or JAK3−/− mice, have defect in lymphoid
organ development (Cao et al. 1995; Park et al. 1995). Additionally, both IL-7 an IL-15 utilize the
common y chain receptor for their signaling and patients with severe combined immunodeficiency
(SCID), who carry mutations of IL2RG or JAK3, are deficient in ILC subsets; further demon­
strating the importance of JAK-STAT pathway in ILC development and homeostasis (Vély et al.
2016). IL-7 and IL-15 signal through JAK3–STAT5 via two STAT5 isoforms including STAT5A
and STAT5B. STAT5B has an important role in regulating development and activation of ILCs
(Imada et al. 1998; Villarino et al. 2017b).
Both IL-12-STAT4 and Type I IFN-STAT1 signaling are important for NK cell development,
though the role of the latter remains controversial (Nguyen et al. 2000; Thierfelder et al. 1996). In
particular, STAT4 is indispensable for the differentiation and proliferation of specific subsets of NK
cells (Madera et al. 2018). In addition, ILC1 cells require IL-12-STAT4 activity to promote their pro-
inflammatory functions against pathogens, especially tissue-resident ILC (Weizman et al. 2017).
ILC2 cells require transcription factors RORα and GATA3 for their development, and secrete
IL-5, IL-9, and IL-13 that are indispensable in immune response to parasitic worm infections and
allergic inflammation (Walker and McKenzie 2013). Although IL-4/STAT6 signaling does not
play an essential role for the differentiation of ILC2, it is still important for ILC2 function in
specific inflammatory conditions. For example, reduced IL-13-producing ILC2 were observed in
STAT6-deficient mice during Nippostrongylus brasiliensis infection and in a mouse model of lung
inflammation induced by the fungal allergen Alternaria (Doherty et al. 2012; Liang et al. 2011).
ILC3 cells require Rorγt and Ahr for their development and production of IL-17A, IL-17F, and
IL-22 cytokines that regulate intestinal homeostasis, infection, and inflammation (Melo-Gonzalez
and Hepworth 2017). IL-23/STAT3 signaling is required to promote ILC3 functions (Geremia
et al. 2011; Takatori et al. 2009). Similar to T helper cell development, a combination of multiple
STATs are involved in the development of ILC subsets (Stabile et al. 2018). While STAT3 is
important for ILC3 functions, it is also essential for NK cell effector function (Zhu et al. 2014). In
addition, IL-23 activates STAT4 and is required for the development of the IFN-γ-producing
NCR+ILC3 cell population (Mikami et al. 2018).

13.4 Mutations of JAKs and STATs Associated with Primary Immunodeficiency


Given the important role of JAK-STAT, defects or mutation of JAK-STAT have been linked to
several disorders of primary immunodeficiency. JAK1 mutations are found in 20% of down
syndrome-associated B-ALL and 9% of patients with hepatitis B-associated hepatocellular
carcinoma (Hammarén et al. 2018). Two homozygous missense germline mutations, including
P733L and P832S, of JAK1 have been identified (Eletto et al. 2016). Patients with the mutations
have diminished expression of IFN-target genes due to reduced JAK1–STAT signaling, which
results in enhanced susceptibility to atypical mycobacterial infection and metastatic bladder
carcinoma (Eletto et al. 2016). In addition, frameshift mutations in JAK1 have been identified
that cause increased mutation burden and microsatellite instability (Albacker et al. 2017). Gene
expression analysis from biopsy samples of patients with JAK1 frameshift mutations reveals
defects in multiple interferon and anti-tumor immune response signatures (Stelloo et al. 2016).
246 JAK-STAT Signaling in Diseases

Altogether, JAK1 mutations lead to the loss of multiple IFN-mediated immune responses against
tumor and bacterial infections.
JAK2 has a high mutation frequency of 38.7% (Villarino et al. 2017a). Erythropoietin and
thrombopoietin-mediated JAK2 signaling pathway is important for the expansion of erythrocytes
and megakaryocytes (Ihle and Gilliland 2007). The most common JAK2 mutation is V617F of
pseudokinase domain that is responsible for its inhibitory function (Nielsen et al. 2011).
JAK3 is commonly expressed in T cells, NK cells, and intestinal epithelial cells (Kawamura
et al. 1994; Musso et al. 1995). JAK3 signaling involves receptors containing the common gamma
chain (γc) of type I cytokine receptor family including IL-2R, IL-4R, IL-7R, IL-9R, IL-15R, and
IL-21R (Babon et al. 2014). Because JAK3 signaling is associated with γc, its mutation has been
found to be similar to X-link severe combined immunodeficiency (SCID) underlying by γc
mutation (Pesu et al. 2005). The diagnosis of SCID is generally based on recurrent severe
infection, diarrhea, and atopic dermatitis (Pesu et al. 2005). Patients with JAK3 or γc mutation
are unable to mount a proper immune responses due to nonfunctional T cells, defective B cell
activation, and impaired immunoglobulin production (Pesu et al. 2005).
Tyrosine kinase 2 (TYK2) signals through the receptors of type I IFN, IL-6, IL-10, IL-12, and IL­
23 (Strobl 2011). A frameshift at codon 90 disrupts the translation of TYK2 protein and results in
mutation (Nemoto et al. 2018). TYK2 mutation was discovered based on a patient with symptoms of
hyper-IgE syndrome (HIES), that developed infection with bacteria including Staphylococcus and
mycobacteria and/or viruses (Minegishi et al. 2006). In a mild case, a recessive partial TYK2
deficiency with heterozygous mutations (c.209_212delGCTT/c.691C > T, p.Cys70Serfs*21/p.
Arg231Trp) was found in patients with T-cell lymphopenia who developed Epstein–Barr virus (EBV)­
associated B-cell lymphoma (Nemoto et al. 2018). Unlike the case of complete TYK2 mutation,
patients with partial TYK2 deficiency have impaired T cell responses to IL-23 (Nemoto et al. 2018).
Genome-wide association study (GWAS) studies show an association between polymorphisms of
TYK2 with various autoimmune diseases including inflammatory bowel disease, systemic lupus
erythematosus, multiple sclerosis, psoriasis, and rheumatoid arthritis (Dendrou et al. 2016).
STAT1 mutations include both gain-of-function (GOF) and loss-of-function (LOF) mutations
(Villarino et al. 2017a). GOF mutation in STAT1 is associated with chronic mucocutaneous
candidiasis (CMC) characterized by persistent Candida albicans infections of skin, nails, and
mucosa (Carey et al. 2018). Moreover, STAT1 GOF mutations have been linked with autoimmu­
nity and IPEX-like syndrome (Uzel et al. 2013). The mechanisms underlying these abnormalities
are due to enhanced IFN-γ signaling, thus inhibiting IL-17 production and increasing suscept­
ibility to fungal infections (Uzel et al. 2013). In contrast, patients with LOF mutation in STAT1,
have increased susceptibility to mycobacterial infections due to a reduction in type II IFN
signaling (Boisson-Dupuis et al. 2012). Another variant of STAT1 LOF have impaired type
I and type II IFN signaling, thus causing reduced anti-viral response (O’Shea et al. 2013).
Type I IFN signaling through STAT2 is important for immunity against viral infection (Morrison
and García-Sastre 2014). Patients with STAT2 mutations have been identified including STAT2
c.381+5 G>C and c.1836 C>A (Hambleton et al. 2013; Shahni et al. 2015). STAT2 c.381+5 G>C
results in a splice mutation, and patients with this mutation are susceptible to severe viral infection
(Hambleton et al. 2013). In the latter case, STAT2 mutation was identified using whole exome
sequencing from patient with severe neurological deterioration following viral infection (Shahni et al.
2015). Patients with STAT2 c.1836 C>A have elongated mitochondria due to inactivation of fission
protein DRP1 (encoded by DNM1L) (Shahni et al. 2015). Thus, STAT2 mutation can lead to defects
in mitochondrial dynamics and could be the cause of neurological disease (Hambleton et al. 2013).
IL-6 and other gp130 cytokines signal through STAT3 to induce gene expression and are critical
in various cellular processes including cell growth and apoptosis (Hillmer et al. 2016). Dominant
negative mutation of STAT3 is associated with hyperimmunoglobulin E syndrome (HIES, or Job’s
syndrome) that is characterized by recurring and severe cutaneous and sinopulmonary bacterial
infections, increased serum immunoglobulin E (IgE), and connective tissue abnormalities (Holland
et al. 2009). Because IL-6-STAT3 signaling plays a key role in the development of Th17 cells that
secrete IL-17 and IL-22 cytokines to mediate protective immune responses against bacterial and
Immunodeficiency 247

fungal infections at barrier sites, mutations in STAT3 signaling are almost always Th17/IL-17­
related defects (Foley 2007; Weaver et al. 2013; Zindl et al. 2013). Patients with STAT3 mutation
often experience atopic dermatitis, staphylococcal skin abscesses, and mucocutaneous candidiasis
typical of the disease (Mogensen 2013). STAT3 is also required for CD8 memory T cells; thus,
impaired STAT3 function results in fragile immune response against viral infections (Siegel et al.
2011). In contrast to STAT3 LOF that causes immunodeficiency, GOF mutations in STAT3 result
in defective STAT1 and STAT5 phosphorylation and regulatory T cell development, and are
associated with autoimmune diseases (Milner et al. 2015; Nabhani et al. 2017). Patients with
STAT3 GOF were diagnosed with lymphadenopathy, autoimmune cytopenias, multi-organ auto­
immunity, and infections (Milner et al. 2015).
IL-12 activation of STAT4 is important for the development of Th1 and Th17 cell differentia­
tion, monocyte activation, and IFN-γ production (Athie-Morales et al. 2004). Although germline
GOF and LOF in STAT4 have not been reported, a number of GWAS studies have identified an
association of STAT4 variants with autoimmune diseases including inflammatory bowel disease,
systemic lupus erythematosus, rheumatoid arthritis, primary biliary cirrhosis, Sjögren′s syndrome,
and systemic sclerosis (Jiang et al. 2013; Korman et al. 2008; Remmers et al. 2007).
STAT5 exists in two highly identical variants, STAT5A and STAT5B, at the amino acid level
(Heltemes-Harris et al. 2011; Rani and Murphy 2016). STAT5A and STAT5B expression has been
detected in mammary tissue and muscle/liver, respectively (Mui et al. 1995; Rani and Murphy
2016). Both STAT5A and STAT5B are activated by various cytokines and hormones such as
growth hormone, erythropoietin, prolactin, and interleukins, and are essential for the develop­
ment and homeostasis of lymphoid cells (Rani and Murphy 2016). Because of the importance of
STAT5 in growth hormone signaling, defect in STAT5 function especially autosomal recessive
STAT5B mutations, can result in dwarfism, immunodeficiency, and autoimmunity (Hwa et al.
2011; Klammt et al. 2018; Kofoed et al. 2003). In addition, STAT5B mutations have been linked
with increased susceptibility to bacterial and viral infections, severe eczema, immunodeficiency,
pulmonary inflammation, and immunodeficiency (Leonard and O’Shea 1998).
STAT6 is activated in response to the cytokine IL-4, and is important in Th2 cell development
and anti-apoptotic activity (Kaplan et al. 1996). Although GOF and LOF in STAT6 have not
been reported, GWAS studies have discovered STAT6 variants that are associated with atopy,
asthma, and eosinophilic esophagitis (Sleiman et al. 2014). Additionally, a novel STAT6 mutation
hotspot (p.419D/G, p.419D/A, and p.419D/H) is responsible for 11% of follicular lymphoma (FL),
a common non-Hodgkin lymphoma (Yildiz et al. 2015).

13.5 Therapeutic Targeting of JAK-STAT Signaling


Because of the importance of JAK-STAT signaling in autoimmune disease and malignancy,
therapeutic approaches targeting the pathway have been intensively developed (Figure 13.3).
One of the approaches is to target the upstream activators including recombinant cytokines,
engineered cytokines, agonist cytokine–antibody complexes, and agonist cytokine–antibody
fusions that have had some success in animal models and clinical trials (O’Shea et al. 2015). For
example, Type I IFN has been used to treat hepatitis C and multiple sclerosis (Reder and Feng
2014; Rong and Perelson 2010). Low dose of IL-2 has been used to target certain autoimmune
diseases (Collison 2018). Moreover, FDA-approved cytokine- and receptor-blocking antibodies,
such as anti-IL2Rα, anti-IL-6, anti-IL-6R, anti-IL-12, and anti-IL-23, have shown promising
results in immunotherapies (Niederreiter et al. 2013; Rider et al. 2016). Given the vital role of the
kinase domain for JAK activity, it has become an attractive target for therapeutic approaches
(O’Shea et al. 2015). The first FDA-approved JAK inhibitor (JAKinibs) is Ruxolitinib that targets
both JAK1 and JAK2 (Mesa 2010). It has been used for treatment of primary myelofibrosis, and
has some encouraging success in psoriasis and in rheumatoid arthritis (RA) (Mesa 2010).
Tofacitinib was developed to block JAK3, that is used for treatment of rheumatoid arthritis,
248 JAK-STAT Signaling in Diseases

FIGURE 13.3 Current therapeutic approaches targeting JAK-STAT. (1, black) Inhibitors target upstream activators
including recombinant cytokines, engineered cytokines, agonist cytokine–antibody complexes, and agonist cytokine–
antibody fusions (2, blue) JAK inhibitors (JAKinibs) and receptor tyrosine kinase (RTK) inhibitors (3, red) STAT
inhibitors

psoriasis, IBD, transplant rejection, juvenile arthritis, and many others (Ellis et al. 2015; Furumoto
and Gadina 2013). JAK3 was thought to be an attractive therapeutic target due to its essential role
in lymphocytes, and that mutations in JAK3 does not affect non-immunologic organs or tissues
(Babon et al. 2014). However, tofacitinib also targets, to a lesser extent, JAK1 and JAK2, thus
causing adverse effect on patients. In fact, mild anemia with neutropenia and increased malignancy
risk were observed in patients treated with tofacitinib (Clark et al. 2014; Riese et al. 2010). Thus,
further studies will be needed to clarify the risk of tofacitinib during the course of treatment.
Another JAK inhibitor, oclacitinib, does not target a specific JAK, and this has raised the question
of the drug’s effectiveness and side effects (Gonzales et al. 2014). However, it has shown promising
results in treating allergic skin disease and could be a potential therapeutic approach to treat
immune-mediated dermatologic conditions in patients (Damsky and King 2017). Other JAK
inhibitors have been developed for various diseases such as asthma, malignancies, and autoimmune
conditions but are currently at different stages of clinical trials (O’Shea et al. 2015). Baricitinib and
momelotinib inhibit both JAK1 and JAK2, and have shown encouraging results in the treatment of
autoinflammatory diseases and myelofibrosis, respectively (Fleischmann et al. 2017; Tefferi et al.
2018). Although the use of JAK inhibitors has yielded some successes, there are remaining issues
that need to be addressed and rigorously interrogated, including the development of more selective
JAK inhibitors and the optimal dosing during treatment.
In contrast to JAK inhibitors that specifically target kinase domains, STAT proteins do not
have catalytic activity and pose more of a challenge for therapeutic targets. STAT inhibitors have
been developed by blocking phosphorylation and thus preventing SH2-mediated binding to
receptors and dimerization or inhibiting STAT DNA-binding activity. Several strategies have
Immunodeficiency 249

been developed to target STATs including oligonucleotide-based STAT inhibitors, small molecule
inhibitors, and STAT-binding intrabodies (Howell et al. 2018; Koo et al. 2014; Sen and Grandis
2014). These STAT inhibitors are still at various stages of the development (O’Shea et al. 2015).
For example, oligonucleotide-based STAT inhibitors are being examined in several malignancies
(Fagard et al. 2014). OPB-51,062 and OPB-31,121.6 have been developed to target STAT3 and
have shown promising results in animal models of autoimmune disease (O’Shea et al. 2015).

13.6 Concluding Remarks


The JAK-STAT represents an ancient signaling pathway that regulates cellular communication
and gene expression. Mutations in the JAK-STAT pathway are frequently associated with primary
immunodeficiencies. Because JAK-STAT plays an important role in the development and home­
ostasis of T and ILC cells, defects in any of its signaling components results in severe immune
dysregulation; therefore, therapeutic targeting of JAK-STAT is challenging due to the requirement
of high specificity. Recent studies show the involvement of JAK-STAT in cell biology expanding
beyond immunology and cancer biology and emerging to the field of stem cell biology and
metabolomics. Thus, JAK-STAT signaling pathway acts as a central hub for the regulation of
molecular and cellular functions. As a result, it is required to interrogate in depth the regulation
of JAK-STAT pathway and its role in regulating chromatin biology and cellular differentiation in
a cell-type specific manner. In addition, the precise mechanisms of how JAK-STAT inhibitors
work in different clinical settings need to be investigated in detail. Thus, a better understanding of
the molecular function of JAK-STAT using advance technology for genetic manipulation, such as
CRISPR-Cas9, would provide tremendous mechanistic insights for improved diagnosis and
therapeutic options for JAK-STAT-related diseases.

REFERENCES
Albacker, Lee A, Jeremy Wu, Peter Smith, Markus Warmuth, Philip J Stephens, Ping Zhu, Lihua Yu, and
Juliann Chmielecki. 2017. Loss of Function JAK1 Mutations Occur at High Frequency in Cancers
with Microsatellite Instability and are Suggestive of Immune Evasion. PLoS One 12 (11):e0176181.
Athie-Morales, Veronica, Hermelijn H Smits, Doreen A Cantrell, and Catharien M U Hilkens. 2004.
Sustained IL-12 Signaling Is Required for Th1 Development. J Immunol 172 (1):61–9.
Au-Yeung, Nancy, Roli Mandhana, and Curt M Horvath. 2014. Transcriptional Regulation by STAT1
and STAT2 in the Interferon JAK–STAT Pathway. JAKStat 2 (3):e23931.
Babon, Jeffrey J, Isabelle S Lucet, James M Murphy, Nicos A Nicola, and Leila N Varghese. 2014. “the
Molecular Regulation of Janus Kinase (JAK) Activation. Biochemical J 462 (1):1–13.
Becher, Burkhard, and Benjamin M Segal. 2011. TH17 Cytokines in Autoimmune Neuro-Inflammation.
Curr Opinion Immunol 23 (6):707–12.
Begitt, Andreas, Mathias Droescher, Klaus-Peter Knobeloch, and Uwe Vinkemeier. 2011. “SUMO
Conjugation of STAT1 Protects Cells from Hyperresponsiveness to IFNγ. Blood 118 (4):1002–7.
Bluestone, Jeffrey A, and Qizhi Tang. 2005. How Do CD4+CD25+ Regulatory T Cells Control
Autoimmunity? Curr Opinion in Immunol 17 (6):638–42.
Böhmer, Frank-D, and Karlheinz Friedrich. 2014. Protein Tyrosine Phosphatases as Wardens of STAT
Signaling. JAKStat 3 (1):e28087.
Boisson-Dupuis, Stephanie, Xiao-Fei Kong, Satoshi Okada, Sophie Cypowyj, Anne Puel, Laurent Abel,
and Jean-Laurent Casanova. 2012. Inborn Errors of Human STAT1: Allelic Heterogeneity Governs
the Diversity of Immunological and Infectious Phenotypes. Curr Opinion in Immunol 24 (4):364–78.
Bousoik, Emira, and Hamidreza Montazeri Aliabadi. 2018. Do We Know Jack’ about JAK? a Closer
Look at JAK/STAT Signaling Pathway. Front in Oncol 8:338.
Cao, X, E W Shores, J Hu-Li, M R Anver, B L Kelsall, S M Russell, J Drago, M Noguchi, A Grinberg,
and E T Bloom. 1995. Defective Lymphoid Development in Mice Lacking Expression of the
Common Cytokine Receptor Gamma Chain. Immunity 2 (3):223–38.
250 JAK-STAT Signaling in Diseases

Carey, Barbara, Jonathan Lambourne, Stephen Porter, and Tim Hodgson. 2018. Chronic Mucocutaneous
Candidiasis Due to Gain-of-Function Mutation in STAT1. Oral Diseases 17 (4). (10.1111):399.
Ciofani, Maria, Aviv Madar, Carolina Galan, MacLean Sellars, Kieran Mace, Florencia Pauli,
Ashish Agarwal et al. 2012. A Validated Regulatory Network for Th17 Cell Specification. Cell 151
(2):289–303.
Clark, James D, Mark E Flanagan, and Jean-Baptiste Telliez. 2014. Discovery and Development of Janus
Kinase (JAK) Inhibitors for Inflammatory Diseases. J Med Chem 57 (12):5023–38.
Collison, Joanna. 2018. Low-Dose IL-2 Therapy for Autoimmune Diseases. Nat Rev Rheumatol 15 (1):2.
Croker, Ben A, Hiu Kiu, and Sandra E Nicholson. 2008. SOCS Regulation of the JAK/STAT Signalling
Pathway. Semin Cell Dev Biol 19 (4):414–22.
Crotty, Shane. 2011. Follicular Helper CD4 T Cells (T FH). Annu Rev Immunol 29 (1):621–63.
Damsky, William, and Brett A King. 2017. JAK Inhibitors in Dermatology: The Promise of a New Drug
Class. J Am Acad Dermatol 76 (4):736–44.
Darnell, J E, Jr. 1997. STATs and Gene Regulation. Science 277 (5332):1630–35.
Dendrou, Calliope A, Adrian Cortes, Lydia Shipman, Hayley G Evans, Kathrine E Attfield, Luke Jostins,
Thomas Barber et al. 2016. Resolving TYK2 Locus Genotype-to-Phenotype Differences in
Autoimmunity. Sci Transl Med 8 (363):363ra149.
Doherty, Taylor A, Naseem Khorram, Jinny E Chang, Hee-Kyoo Kim, Peter Rosenthal, Michael Croft,
and David H Broide. 2012. STAT6 Regulates Natural Helper Cell Proliferation during Lung
Inflammation Initiated by Alternaria. Am J Physiol Lung Cell Mol Physiol 303 (7):L577–88.
Eletto, Davide, Siobhan O Burns, Ivan Angulo, Vincent Plagnol, Kimberly C Gilmour,
Frances Henriquez, James Curtis et al. 2016. Biallelic JAK1 Mutations in Immunodeficient Patient
with Mycobacterial Infection. Nat Commun 7 (1):13992.
Ellis, Martin H, Noa Lavi, Elena Mishchenko, Najib Dally, David Lavie, Anna Courevitch, Odit Gutwein
et al. 2015. Ruxolitinib Treatment for Myelofibrosis: Efficacy and Tolerability in Routine Practice.
Leukemia Research 39 (11):1154–58.
Fagard, Remi, Valeri Metelev, Inès Souissi, and Fanny Baran-Marszak. 2014. STAT3 Inhibitors for
Cancer Therapy. JAKStat 2 (1):e22882.
Fleischmann, Roy, Jahangir Alam, Vipin Arora, John Bradley, Douglas E Schlichting, David Muram,
and Josef S Smolen. 2017. Safety and Efficacy of Baricitinib in Elderly Patients with Rheumatoid
Arthritis. RMD Open 3 (2):e000546.
Foley, J F. 2007. STAT3 Regulates the Generation of Th17 Cells. Science’s STKE 2007 (380):tw113.
Furumoto, Yasuko, and Massimo Gadina. 2013. The Arrival of JAK Inhibitors: Advancing the Treatment
of Immune and Hematologic Disorders. BioDrugs 27 (5):431–8.
Geremia, Alessandra, Carolina V Arancibia-Cárcamo, Myles P P Fleming, Nigel Rust, Baljit Singh, Neil
J Mortensen, Simon P L Travis, and Fiona Powrie. 2011. IL-23–Responsive Innate Lymphoid Cells
Are Increased in Inflammatory Bowel Disease. J Exp Med 208 (6):1127–33.
G Ler, M L, J D Gorham, C S Hsieh, A J Mackey, R G Steen, W F Dietrich, and K M Murphy. 1996.
Genetic Susceptibility to Leishmania: IL-12 Responsiveness in TH1 Cell Development. Science 271
(5251):984–7.
Gonzales, A J, J W Bowman, G J Fici, M Zhang, D W Mann, and M Mitton-Fry. 2014. Oclacitinib
(APOQUEL ®) Is a Novel Janus Kinase Inhibitor with Activity against Cytokines Involved in
Allergy. J Vet Pharmacol Ther 37 (4):317–24.
Goswami, Ritobrata, Rukhsana Jabeen, Ryoji Yagi, Duy Pham, Jinfang Zhu, Shreevrat Goenka, and Mark
H Kaplan. 2012. “STAT6-dependent Regulation of Th9 Development. J Immunol 188 (3):968–75.
Hambleton, Sophie, Stephen Goodbourn, Dan F Young, Paul Dickinson, Siti M B Mohamad,
Manoj Valappil, Naomi McGovern et al. 2013. STAT2 Deficiency and Susceptibility to Viral Illness
in Humans. Proc Natl Acad Sci U S A 110 (8):3053–8.
Hammarén, Henrik M, Anniina T Virtanen, Juuli Raivola, and Olli Silvennoinen. 2018. The Regulation
of JAKs in Cytokine Signaling and Its Breakdown in Disease. Cytokine 118:48–63 .
Heltemes-Harris, Lynn M, Mark J L Willette, Kieng B Vang, and Michael A Farrar. 2011. The Role of
STAT5 in the Development, Function, and Transformation of B and T Lymphocytes. Ann
N Y Acad Sci 1217 (1):18.
Hillmer, Emily J, Huiyuan Zhang, Haiyan S Li, and Stephanie S Watowich. 2016. STAT3 Signaling in
Immunity. Cytokine Growth Factor Rev 31:1–15.
Immunodeficiency 251

Holland, Steven M, Frank R DeLeo, Houda Z Elloumi, Amy P Hsu, Gulbu Uzel, Nina Brodsky,
Alexandra F Freeman et al. 2009. STAT3 Mutations in the Hyper-IgE Syndrome. Dx.Doi.org 357
(16):1608–19.
Howell, Michael D, Carolyn Fitzsimons, and Paul A Smith. 2018. JAK/STAT Inhibitors and Other Small
Molecule Cytokine Antagonists for the Treatment of Allergic Disease. Ann Allergy Asthma
Immunol 120 (4):367–75.
Hwa, Vivian, Kari Nadeau, Jan M Wit, and Ron G Rosenfeld. 2011. STAT5b Deficiency: Lessons From
STAT5b Gene Mutations. Best Pract Res Clin Endocrinol Metab 25 (1):61–75.
Ihle, James N, and D Gary Gilliland. 2007. JAK2: Normal Function and Role in Hematopoietic
Disorders. Curr Opin Genet Dev 17 (1):8–14.
Imada, Kazunori, Eda T Bloom, Hiroshi Nakajima, Judith A Horvath-Arcidiacono, Garry B Udy, Helen
W Davey, and Warren J Leonard. 1998. Stat5b Is Essential for Natural Killer Cell–Mediated
Proliferation and Cytolytic Activity. J Exp Med 188 (11):2067–74.
Jiang, De-Ke, Jielin Sun, Guangwen Cao, Yao Liu, Dongxin Lin, Yu-Zhen Gao, Wei-Hua Ren et al.
2013. Genetic Variants in STAT4 and HLA-DQ Genes Confer Risk of Hepatitis B Virus-Related
Hepatocellular Carcinoma. Nat Genet 45 (1):72–5.
John, S, U Vinkemeier, E Soldaini, J E Darnell, and W J Leonard. 1999. The Significance of Tetrameriza­
tion in Promoter Recruitment by Stat5. Mol Cell Biol 19 (3):1910–18.
Kaplan, Mark H, Matthew M Hufford, and Matthew R Olson. 2015. The Development and in Vivo
Function of T Helper 9 Cells. Nat Rev Immunol 15 (5):295–307.
Kaplan, Mark H, Ulrike Schindler, Stephen T Smiley, and Michael J Grusby. 1996. Stat6 Is Required for
Mediating Responses to IL-4 and for the Development of Th2 Cells. Immunity 4 (3):313–19.
Kawamura, M, D W McVicar, J A Johnston, T B Blake, Y Q Chen, B K Lal, A R Lloyd, D J Kelvin,
J E Staples, and J R Ortaldo. 1994. Molecular Cloning of L-JAK, a Janus Family Protein-Tyrosine
Kinase Expressed in Natural Killer Cells and Activated Leukocytes. Proc Natl Acad Sci U S A 91
(14):6374–8.
Klammt, Jürgen, David Neumann, Evelien F Gevers, Shayne F Andrew, I David Schwartz,
Denise Rockstroh, Roberto Colombo et al. 2018. Dominant-Negative STAT5B Mutations Cause
Growth Hormone Insensitivity with Short Stature and Mild Immune Dysregulation. Nat Commun
9 (1):472.
Kofoed, Eric M, Vivian Hwa, Brian Little, Katie A Woods, Caroline K Buckway, Junko Tsubaki,
Katherine L Pratt et al. 2003. Growth Hormone Insensitivity Associated with a
STAT5bMutation. New Engl J Med 349 (12):1139–47.
Koo, Mi Young, Jiyoung Park, Jung Mi Lim, Sei Yoon Joo, Seung-Pil Shin, Hyun Bo Shim,
Junho Chung, Dongmin Kang, Hyun Ae Woo, and Sue Goo Rhee. 2014. Selective Inhibition of
the Function of Tyrosine-Phosphorylated STAT3 with a Phosphorylation Site-Specific Intrabody.”.
Proc Natl Acad Sci U S A 111 (17):6269–74.
Korman, B D, M I Alba, J M Le, I Alevizos, J A Smith, N P Nikolov, D L Kastner, E F Remmers, and
G G Illei. 2008. Variant Form of STAT4 Is Associated with Primary Sjögren’s Syndrome. Genes
Immun 9 (3):267–70.
Leonard, Warren J. 2002. The JAK–STAT Pathway. In Hormone Signaling 17:103–120.
Leonard, Warren J, and John J O’Shea. 1998. JAKS and STATS: Biological Implications. Annu Rev
Immunol 16 (1):293–322.
Liang, Hong-Erh, R Lee Reinhardt, Jennifer K Bando, Brandon M Sullivan, I-Cheng Ho, and
Richard M Locksley. 2011. Divergent Expression Patterns of IL-4 and IL-13 Define Unique
Functions in Allergic Immunity. Nat Immunol 13 (1):58–66.
Liao, Wei, Dustin E Schones, Jangsuk Oh, Yongzhi Cui, Kairong Cui, Tae-Young Roh, Keji Zhao, and
Warren J Leonard. 2008. Priming for T Helper Type 2 Differentiation by Interleukin 2-Mediated
Induction of Interleukin 4 Receptor Alpha-Chain Expression. Nat Immunol 9 (11):1288–96. doi:
10.1038/ni.1656.
Lim, Cheh Peng, and Xinmin Cao. 2006. Structure, Function, and Regulation of STAT Proteins. Mol
BioSyst 2 (11):536–50.
Lin, Jian-Xin, Ning Du, Peng Li, Majid Kazemian, Tesfay Gebregiorgis, Rosanne Spolski, and Warren
J Leonard. 2017. Critical Functions for STAT5 Tetramers in the Maturation and Survival of
Natural Killer Cells. Nat Commun 8 (1):293.
252 JAK-STAT Signaling in Diseases

Lin, Jian-Xin, Peng Li, Delong Liu, Hyun Tak Jin, Jianping He, Mohammed Ata Ur Rasheed,
Yrina Rochman et al. 2012. Critical Role of STAT5 Transcription Factor Tetramerization for
Cytokine Responses and Normal Immune Function. Immunity 36 (4):586–99.
Lodolce, James P, David L Boone, Sophia Chai, Rachel E Swain, Themistocles Dassopoulos,
Shanthi Trettin, and Averil Ma. 1998. IL-15 Receptor Maintains Lymphoid Homeostasis by
Supporting Lymphocyte Homing and Proliferation. Immunity 9 (5):669–76.
Lu, Dan, Liang Liu, Xin Ji, Yanan Gao, Xi Chen, Yu Liu, Yang Liu et al. 2015. The Phosphatase DUSP2
Controls the Activity of the Transcription Activator STAT3 and Regulates TH17 Differentiation.
Nat Immunol 16 (12):1263–73.
Ma, Da, Hua Huang, and Zan Huang. 2010. STAT1 Signaling Is Required for Optimal Th1 Cell
Differentiation in Mice. Chin Sci Bull 55 (11):1032–40. doi: 10.1007/s11434-010-0030-9.
Madera, Sharline, Clair D Geary, Colleen M Lau, Olga Pikovskaya, Steven L Reiner, and Joseph C Sun.
2018. Cutting Edge: Divergent Requirement of T-Box Transcription Factors in Effector and
Memory NK Cells. J Immunol 200 (6):1977–81.
Mahmud, Shawn A, Luke S Manlove, and Michael A Farrar. 2013. Interleukin-2 and STAT5 in
Regulatory T Cell Development and Function. JAKStat 2 (1):e23154.
Maier, Elisabeth, Albert Duschl, and Jutta Horejs-Hoeck. 2012. STAT6-Dependent and -Independent
Mechanisms in Th2 Polarization. Eur J Immunol 42 (11):2827–33.
Meier, Jeremy A, and Andrew C Larner. 2014. Toward a New STATe: The Role of STATs in Mitochon­
drial Function. Semin Immunol 26 (1):20–8.
Melo-Gonzalez, Felipe, and Matthew R Hepworth. 2017. Functional and Phenotypic Heterogeneity of
Group 3 Innate Lymphoid Cells. Immunology 150 (3):265–75.
Mesa, Ruben A. 2010. Ruxolitinib, a Selective JAK1 and JAK2 Inhibitor for the Treatment of
Myeloproliferative Neoplasms and Psoriasis. IDrugs 13 (6):394–403.
Meyer, Thomas, Lisa Hendry, Andreas Begitt, Susan John, and Uwe Vinkemeier. 2004. A Single Residue
Modulates Tyrosine Dephosphorylation, Oligomerization, and Nuclear Accumulation of Stat
Transcription Factors. J Biol Chem 279 (18):18998–19007.
Mikami, Yohei, Gianluca Scarno, Beatrice Zitti, Han-Yu Shih, Yuka Kanno, Angela Santoni, John J O’Shea,
and Giuseppe Sciumè. 2018. NCR+ ILC3 Maintain Larger STAT4 Reservoir via T-BET to Regulate
Type 1 Features upon IL-23 Stimulation in Mice. Eur J Immunol 48 (7):1174–80.
Milner, Joshua D, Tiphanie P Vogel, Lisa Forbes, Chi A Ma, Asbjørg Stray-Pedersen, Julie E Niemela,
Jonathan J Lyons et al. 2015. Early-Onset Lymphoproliferation and Autoimmunity Caused by
Germline STAT3 Gain-of-Function Mutations. Blood 125 (4):591–9.
Minegishi, Yoshiyuki, Masako Saito, Tomohiro Morio, Ken Watanabe, Kazunaga Agematsu,
Shigeru Tsuchiya, Hidetoshi Takada et al. 2006. Human Tyrosine Kinase 2 Deficiency Reveals Its
Requisite Roles in Multiple Cytokine Signals Involved in Innate and Acquired Immunity. Immunity
25 (5):745–55.
Mogensen, Trine H. 2013. STAT3 and the Hyper-IgE Syndrome: Clinical Presentation, Genetic Origin,
Pathogenesis, Novel Findings and Remaining Uncertainties. JAKStat 2 (2):e23435.
Morrison, Juliet, and Adolfo García-Sastre. 2014. STAT2 Signaling and Dengue Virus Infection.
JAKStat 3 (1):e27715.
Mowen, Kerri A, Jie Tang, Wei Zhu, Brandon T Schurter, Ke Shuai, Harvey R Herschman, and
Michael David. 2001. Arginine Methylation of STAT1 Modulates IFNα/Β-Induced Transcription.
Cell 104 (5):731–41.
Mui, A L, H Wakao, A M O’Farrell, N Harada, and A Miyajima. 1995. Interleukin-3,
Granulocyte-Macrophage Colony Stimulating Factor and Interleukin-5 Transduce Signals through
Two STAT5 Homologs. Embo J 14 (6):1166–75.
Murray, P J. 2007. The JAK–STAT Signaling Pathway: Input and Output Integration. J Immunol 178
(5):2623–29.
Musso, T, J A Johnston, D Linnekin, L Varesio, T K Rowe, J J O’Shea, and D W McVicar. 1995.
Regulation of JAK3 Expression in Human Monocytes: Phosphorylation in Response to Interleukins
2, 4, and 7. J Exp Med 181 (4):1425–31.
Nabhani, Schafiq, Cyrill Schipp, Hagit Miskin, Carina Levin, Sergey Postovsky, Tal Dujovny, Ariel Koren
et al. 2017. STAT3 Gain-of-Function Mutations Associated with Autoimmune Lymphoproliferative
Immunodeficiency 253

Syndrome like Disease Deregulate Lymphocyte Apoptosis and Can Be Targeted by BH3 Mimetic
Compounds. Clin Immunol 181:32–42.
Nan, Yuchen, Chunyan Wu, and Yan-Jin Zhang. 2018. Interferon Independent Non-Canonical STAT
Activation and Virus Induced Inflammation. Viruses 10 (4):196.
Nemoto, Michiko, Hiroyoshi Hattori, Naoko Maeda, Nobuhiro Akita, Hideki Muramatsu,
Suzuko Moritani, Tomonori Kawasaki et al. 2018. Compound Heterozygous TYK2 Mutations
Underlie Primary Immunodeficiency with T-Cell Lymphopenia. Sci Rep 8 (1):515.
Nguyen, Khuong B, Leslie P Cousens, Lesley A Doughty, Gary C Pien, Joan E Durbin, and Christine
A Biron. 2000. Interferon Α/B-Mediated Inhibition and Promotion of Interferon Γ: STAT1
Resolves a Paradox. Nat Immunol 1 (1):70–6.
Niederreiter, Lukas, Timon Adolph, and Arthur Kaser. 2013. Anti-IL-12/23 in Crohn’S Disease: Bench
and Bedside. Curr Drug Targets 14 (12):1379–84.
Nielsen, Camilla, Henrik S Birgens, Børge G Nordestgaard, Lasse Kjaer, and Stig E Bojesen. 2011. The
JAK2 V617F Somatic Mutation, Mortality and Cancer Risk in the General Population. Haemato­
logica 96 (3). Haematologica:450–3.
Niu, Guo-Juan, Ji-Dong Xu, Wen-Jie Yuan, Jie-Jie Sun, Ming-Chong Yang, Zhong-Hua He, Xiao-Fan
Zhao, and Jin-Xing Wang. 2018. Protein Inhibitor of Activated STAT (PIAS) Negatively Regulates the
JAK/STAT Pathway by Inhibiting STAT Phosphorylation and Translocation. Front Immunol 9:2392.
O’Shea, John J, Steven M Holland, and Louis M Staudt. 2013. JAKs and STATs in Immunity,
Immunodeficiency, and Cancer. N Engl J Med 368 (2):161–70.
O’Shea, John J, Riitta Lahesmaa, Golnaz Vahedi, Arian Laurence, and Yuka Kanno. 2011. Genomic Views
of STAT Function in CD4+ T Helper Cell Differentiation. Nat Rev Immunol 11 (4):239–50.
O’Shea, John J, and Peter J Murray. 2008. Cytokine Signaling Modules in Inflammatory Responses.
Immunity 28 (4):477–87.
O’Shea, John J, Daniella M Schwartz, Alejandro V Villarino, Massimo Gadina, Iain B McInnes, and
Arian Laurence. 2015. The JAK–STAT Pathway: Impact on Human Disease and Therapeutic
Intervention. Annu Rev Med 66 (1):311–28.
Olson, Matthew R, Felipe Fortino Verdan, Matthew M Hufford, Alexander L Dent, and Mark
H Kaplan. 2016. STAT3 Impairs STAT5 Activation in the Development of IL-9-Secreting T Cells.
J Immunol 196 (8):3297–304.
Palmer, Douglas C, and Nicholas P Restifo. 2009. Suppressors of Cytokine Signaling (SOCS) in T Cell
Differentiation, Maturation, and Function. Trend Immunol 30 (12):592–602.
Park, S Y, K Saijo, T Takahashi, M Osawa, H Arase, N Hirayama, K Miyake, H Nakauchi, T Shirasawa,
and T Saito. 1995. Developmental Defects of Lymphoid Cells in JAK3 Kinase-Deficient Mice.
Immunity 3 (6):771–82.
Passerini, Laura, Sarah E Allan, Manuela Battaglia, Sara Di Nunzio, Alicia N Alstad, Megan K Levings,
Maria G Roncarolo, and Rosa Bacchetta. 2008. STAT5-Signaling Cytokines Regulate the Expression
of FOXP3 in CD4+CD25+ Regulatory T Cells and CD4+CD25- Effector T Cells. Int Immunol 20
(3):421–31.
Pesu, Marko, Fabio Candotti, Matthew Husa, Sigrun R Hofmann, Luigi D Notarangelo, and John
J O’Shea. 2005. JAK3, Severe Combined Immunodeficiency, and a New Class of Immunosuppressive
Drugs. Immunol Rev 203 (1):127–42.
Qi, Qian-Rong, and Zeng-Ming Yang. 2014. Regulation and Function of Signal Transducer and
Activator of Transcription 3. World J Biol Chem 5 (2):231–9.
Ramos, Hilario J, Ann M Davis, Thaddeus C George, and J David Farrar. 2007. IFN-Alpha Is Not
Sufficient to Drive Th1 Development Due to Lack of Stable T-Bet Expression. J Immunol 179
(6):3792–803.
Rani, Aradhana, and John J Murphy. 2016. STAT5 in Cancer and Immunity. J Interferon Cytokine Res 36
(4):226–37.
Ranson, Thomas, Christian A J Vosshenrich, Erwan Corcuff, Odile Richard, Werner Müller, and James
P Di Santo. 2003. IL-15 Is an Essential Mediator of Peripheral NK-Cell Homeostasis. Blood 101
(12):4887–93.
Rawlings, J S. 2004. The JAK/STAT Signaling Pathway. J Cell Sci 117 (8):281–3.
Reder, Anthony T, and Xuan Feng. 2014. How Type I Interferons Work in Multiple Sclerosis and Other
Diseases: Some Unexpected Mechanisms. J Interferon Cytokine Res 34 (8):589–99.
254 JAK-STAT Signaling in Diseases

Remmers, Elaine F, Robert M Plenge, Annette T Lee, Robert R Graham, Geoffrey Hom, Timothy
W Behrens, Paul I W de Bakker et al. 2007. STAT4 and the Risk of Rheumatoid Arthritis and
Systemic Lupus Erythematosus. N Engl J Med 357 (10):977–86.
Rider, Peleg, Yaron Carmi, and Idan Cohen. 2016. Biologics for Targeting Inflammatory Cytokines,
Clinical Uses, and Limitations. Int J Cell Biol 2016 (7):1–11.
Riese, Richard J, Sriram Krishnaswami, and Joel Kremer. 2010. Inhibition of JAK Kinases in Patients
with Rheumatoid Arthritis: Scientific Rationale and Clinical Outcomes. Best Pract Res Clin
Rheumatol 24 (4):513–26.
Rong, Libin, and Alan S Perelson. 2010. Treatment of Hepatitis C Virus Infection with Interferon
and Small Molecule Direct Antivirals: Viral Kinetics and Modeling. Crit Rev Immunol 30
(2):131–48.
Schindler, Christian. 1999. Cytokines and JAK–STAT Signaling. Exp Cell Res 253 (1):7–14.
Seidel, H M, L H Milocco, P Lamb, J E Darnell, R B Stein, and J Rosen. 1995. Spacing of Palindromic
Half Sites as a Determinant of Selective STAT (Signal Transducers and Activators of Transcription)
DNA Binding and Transcriptional Activity. Proc Natl Acad Sci U S A 92 (7):3041–5.
Sen, Malabika, and Jennifer R Grandis. 2014. Nucleic Acid-Based Approaches to STAT Inhibition.
JAKStat 1 (4):285–91.
Shahni, Rojeen, Catherine M Cale, Glenn Anderson, Laura D Osellame, Sophie Hambleton, Thomas
S Jacques, Yehani Wedatilake et al. 2015. Signal Transducer and Activator of Transcription 2
Deficiency Is a Novel Disorder of Mitochondrial Fission. Brain 138:2834–46.
Shuai, K. 2000. Modulation of STAT Signaling by STAT-Interacting Proteins. Oncogene 19 (21):2638–44.
Shuai, Ke, and Bin Liu. 2003. Regulation of JAK–STAT Signalling in the Immune System. Nat Rev
Immunol 3 (11):900–11.
Siegel, Andrea M, Jennifer Heimall, Alexandra F Freeman, Amy P Hsu, Erica Brittain, Jason M Brenchley,
Daniel C Douek et al. 2011. A Critical Role for STAT3 Transcription Factor Signaling in the
Development and Maintenance of Human T Cell Memory. Immunity 35 (5):806–18.
Singh, Rekha. 2011. JAK2-Independent Activation of Stat3 by Intracellular Angiotensin II in Human
Mesangial Cells. J Signal Transduct 2011 (6):257862–10.
Sleiman, Patrick M A, Mei-Lun Wang, Antonella Cianferoni, Seema Aceves, Nirmala Gonsalves,
Kari Nadeau, Albert J Bredenoord, Glenn T Furuta, Jonathan M Spergel, and Hakon Hakonarson.
2014. GWAS Identifies Four Novel Eosinophilic Esophagitis Loci. Nat Commun 5 (1):5593.
Srivastava, Shivani, Meghan A Koch, Marion Pepper, and Daniel J Campbell. 2014. Type I Interferons
Directly Inhibit Regulatory T Cells to Allow Optimal Antiviral T Cell Responses during Acute
LCMV Infection. J Exp Med 211 (5):961–74.
Stabile, Helena, Gianluca Scarno, Cinzia Fionda, Angela Gismondi, Angela Santoni, Massimo Gadina,
and Giuseppe Sciumè. 2018. JAK/STAT Signaling in Regulation of Innate Lymphoid Cells: The
Gods before the Guardians. Immunol Rev 286 (1):148–59.
Stark, George R, and James E Darnell Jr. 2012. The JAK–STAT Pathway at Twenty. Immunity 36
(4):503–14.
Stelloo, Ellen, Marco A Versluis, Hans W Nijman, Marco de Bruyn, Annechien Plat, Elisabeth M Osse,
Reinhardt H van Dijk et al. 2016. Microsatellite Instability Derived JAK1 Frameshift Mutations
are Associated with Tumor Immune Evasion in Endometrioid Endometrial Cancer. Oncotarget 7
(26):39885–93.
Stritesky, Gretta L, Rajarajeswari Muthukrishnan, Sarita Sehra, Ritobrata Goswami, Duy Pham,
Jared Travers, Evelyn T Nguyen, David E Levy, and Mark H Kaplan. 2011. The Transcription
Factor STAT3 Is Required for T Helper 2 Cell Development. Immunity 34 (1):39–49.
Strobl, Birgit. 2011. Tyrosine Kinase 2 (TYK2) in Cytokine Signalling and Host Immunity. Front Biosci
(Landmark Ed) 16 (1):3214–32.
Takatori, Hiroaki, Yuka Kanno, Wendy T Watford, Cristina M Tato, Greta Weiss, Ivaylo I Ivanov, Dan
R Littman, and John J O’Shea. 2009. Lymphoid Tissue Inducer-Like Cells are an Innate Source of
IL-17 and IL-22. J Exp Med 206 (1):35–41.
Tefferi, Ayalew, Daniela Barraco, Terra L Lasho, Sahrish Shah, Kebede H Begna, Aref Al-Kali, William
J Hogan et al. 2018. Momelotinib Therapy for Myelofibrosis: A 7-Year Follow-Up. Blood Cancer J 8
(3):1322.
Immunodeficiency 255

Thierfelder, William E, Jan M van Deursen, Koh Yamamoto, Ralph A Tripp, Sally R Sarawar, Richard
T Carson, Mark Y Sangster et al. 1996. Requirement for Stat4 in Interleukin-12-Mediated
Responses of Natural Killer and T Cells. Nature 382 (6587):171–4.
Thieu, Vivian T, Qing Yu, Hua-Chen Chang, Norman Yeh, Evelyn T Nguyen, Sarita Sehra, and Mark
H Kaplan. 2008. Signal Transducer and Activator of Transcription 4 Is Required for the Transcrip­
tion Factor T-Bet to Promote T Helper 1 Cell-Fate Determination. Immunity 29 (5):679–90.
Ulrich, Benjamin J, Felipe Fortino Verdan, Andrew N J McKenzie, Mark H Kaplan, and Matthew
R Olson. 2017. STAT3 Activation Impairs the Stability of Th9 Cells. J Immunol 198 (6):2302–9.
Ungureanu, Daniela, Sari Vanhatupa, Noora Kotaja, Jie Yang, Saara Aittomaki, Olli A Jänne, Jorma
J Palvimo, and Olli Silvennoinen. 2003. PIAS Proteins Promote SUMO-1 Conjugation to STAT1.
Blood 102 (9):3311–3.
Uzel, Gulbu, Elizabeth P Sampaio, Monica G Lawrence, Amy P Hsu, Mary Hackett, Morna J Dorsey,
Richard J Noel et al. 2013. Dominant Gain-of-Function STAT1 Mutations in FOXP3 Wild-Type
Immune Dysregulation-Polyendocrinopathy-Enteropathy-X-Linked-Like Syndrome. J Allergy Clin
Immunol 131 (6):1611–23.
Vahedi, Golnaz, Hayato Takahashi, Shingo Nakayamada, Hong-wei Sun, Vittorio Sartorelli,
Yuka Kanno, and John J O’Shea. 2012. STATs Shape the Active Enhancer Landscape of T Cell
Populations. Cell 151 (5):981–93.
Van Nguyen, Thang, Pornpimon Angkasekwinai, Hong Dou, Feng-Ming Lin, Long-Sheng Lu, Jinke Cheng,
Y Eugene Chin, Chen Dong, and Edward T H Yeh. 2012. SUMO-Specific Protease 1 Is Critical for
Early Lymphoid Development through Regulation of STAT5 Activation. Mol Cell 45 (2):210–21.
Vély, Frédéric, Vincent Barlogis, Blandine Vallentin, Bénédicte Neven, Christelle Piperoglou,
Mikael Ebbo, Thibaut Perchet et al. 2016. Evidence of Innate Lymphoid Cell Redundancy in
Humans. Nat Immunol 17 (11):1291–9.
Villarino, Alejandro, Linda Hibbert, Linda Lieberman, Emma Wilson, Tak Mak, Hiroki Yoshida, Robert
A Kastelein, Christiaan Saris, and Christopher A Hunter. 2003. The IL-27R (WSX-1) Is Required
to Suppress T Cell Hyperactivity during Infection. Immunity 19 (5):645–55.
Villarino, Alejandro V, Yuka Kanno, John R Ferdinand, and John J O’Shea. 2015. Mechanisms of JAK/
STAT Signaling in Immunity and Disease. J Immunol 194 (1):21–7.
Villarino, Alejandro V, Yuka Kanno, and John J O’Shea. 2017a. Mechanisms and Consequences of JAK–
STAT Signaling in the Immune System. Nat Immunol 18 (4):374–84.
Villarino, Alejandro V, Giuseppe Sciumè, Fred P Davis, Shigeru Iwata, Beatrice Zitti, Gertraud
W Robinson, Lothar Hennighausen, Yuka Kanno, and John J O’Shea. 2017b. Subset- and
Tissue-Defined STAT5 Thresholds Control Homeostasis and Function of Innate Lymphoid Cells.
Journal Exp Med 214 (10):2999–3014.
Vonarbourg, Cedric, and Andreas Diefenbach. 2012. Multifaceted Roles of Interleukin-7 Signaling for
the Development and Function of Innate Lymphoid Cells. Sem Immunol 24 (3):165–74.
Walker, Jennifer A, and Andrew N J McKenzie. 2013. Development and Function of Group 2 Innate
Lymphoid Cells. Curr Opin Immunol 25 (2):148–55.
Walker, Jennifer A, and Andrew N J McKenzie. 2017. TH2 Cell Development and Function. Nat Rev
Immunol 18 (2):121–33.
Weaver, Casey T, Charles O Elson, Lynette A Fouser, and Jay K Kolls. 2013. The Th17 Pathway and
Inflammatory Diseases of the Intestines, Lungs, and Skin. Annu Rev Pathol 8 (1):477–512.
Wehinger, Jens, Fabrice Gouilleux, Bernd Groner, Juergen Finke, Roland Mertelsmann, and Renate
Maria Weber-Nordt. 1999. IL-10 Induces DNA Binding Activity of Three STAT Proteins (stat1,
Stat3, and Stat5) and Their Distinct Combinatorial Assembly in the Promoters of Selected Genes.
FEBS Lett 394 (3):365–70.
Weizman, Orr-El, Nicholas M Adams, Iona S Schuster, Chirag Krishna, Yuri Pritykin, Colleen Lau,
Mariapia A Degli-Esposti, Christina S Leslie, Joseph C Sun, and Timothy E O’Sullivan. 2017. ILC1
Confer Early Host Protection at Initial Sites of Viral Infection. Cell 171 (4):795–808.e12.
Wu, Hao, Markus M Xie, Hong Liu, and Alexander L Dent. 2016. Stat3 Is Important for Follicular
Regulatory T Cell Differentiation. Edited by Derya Unutmaz. PLoS One 11 (5):e0155040.
Wurtz, O. 2004. IL-4-Mediated Inhibition of IFN-Gamma Production by CD4+ T Cells Proceeds by
Several Developmentally Regulated Mechanisms. Int Immunol 16 (3):501–8.
256 JAK-STAT Signaling in Diseases

Yildiz, M, H Li, D Bernard, N A Amin, P Ouillette, S Jones, K Saiya-Cork et al. 2015. Activating STAT6
Mutations in Follicular Lymphoma. Blood 125 (4):668–79.
Yuan, Z L. 2005. Stat3 Dimerization Regulated by Reversible Acetylation of a Single Lysine Residue.
Science 307 (5707):269–73.
Zhang, S, S Fukuda, Y Lee, G Hangoc, S Cooper, R Spolski, W J Leonard, and H E Broxmeyer. 2000.
Essential Role of Signal Transducer and Activator of Transcription (stat)5a but Not Stat5b for
Flt3-Dependent Signaling. J Exp Med 192 (5):719–28.
Zhu, Shiguo, Prasad V Phatarpekar, Cecele J Denman, Vladimir V Senyukov, Srinivas S Somanchi,
Hoainam T Nguyen-Jackson, Emily M Mace et al. 2014. Transcription of the Activating Receptor
NKG2D in Natural Killer Cells Is Regulated by STAT3 Tyrosine Phosphorylation. Blood 124
(3):403–11.
Zhuang, Shougang. 2013. Regulation of STAT Signaling by Acetylation. Cell Signal 25 (9):1924–31.
Zindl, Carlene L, Jen-Feng Lai, Yun Kyung Lee, Craig L Maynard, Stacey N Harbour, Wenjun Ouyang,
David D Chaplin, and Casey T Weaver. 2013. IL-22-Producing Neutrophils Contribute to Anti­
microbial Defense and Restitution of Colonic Epithelial Integrity during Colitis. Proc Natl Acad Sci
U S A 110 (31):12768–73.
14
Targeting JAK-STAT Pathway for Various
Inflammatory Diseases and Viral Infections

Christina Gavegnano and Raymond F. Schinazi


Center for AIDS Research, Laboratory of Biochemical Pharmacology
Department of Pediatrics, Emory University
Atlanta, GA

14.1 Introduction
The (Janus activating kinase) JAK–signal transducer and activator of transcription (STAT)
pathway has been implicated as a key driver in many diseases, including myelofibrosis and
polycythemia vera (overexpression of JAKs), rheumatoid and psoriatic arthritis, inflammatory
bowel disease, ulcerative colitis, uveitis, systemic lupus erythematosus (SLE), and psoriasis.
Previous treatment for these pathologies centered around symptom relief, rather than selective
targeting of the pathway that is responsible for the dysregulated inflammatory response driving
the disease. With the onset of small molecule inhibitors that can selectively and specifically
target key JAKs involved in controlling downstream inflammation and immunoregulatory
cytokines, exploration for their utility across a broad range of diseases has become a rapidly
expanding field.
JAK inhibitors present with unique advantages over non-steroidal anti-inflammatories
(NSAIDs) such as aspirin, or steroidal anti-inflammatories such as prednisone and cortisone
based agents. For example, NSAIDs, including aspirin, target the cyclooxygenase enzyme
pathway, rendering it inactive upon binding; the end result is multifactorial and weakly
selective to COX-1 versus COX-2 (1). Block of COX signaling primarily confers its anti-
inflammatory effects by non-specific events, including block of prostaglandins, which are
paracrine hormones that confer a wide array of chemical functions including temperature
regulation, homeostasis, and inflammation. Further, block of COX-2 has been associated with
severe adverse effects, including cardiac events leading to death; it is thought that excessive
block of COX-2 within the endothelial cells that comprise the microvasculature within the
heart can result in increased thrombosis, leading to heart attack and death (2,3). Additionally,
COX blockade can also result in decreased clotting, which, in turn, can lead to detrimental
long-term side effects systemically (2).
Steroidal anti-inflammatories also present with lack of specificity, and are often associated
with immunosuppression, which can lead to increased opportunistic infections; broad blunt­
ing of T cell, B cell, and antigen-presenting cell responses to pathogens; and increased
malignancies due to blunted ability to sense eliminate potential malignant cells. Together,
the onset of selective, specific inhibitors of the JAK-STAT pathway presents with a unique
opportunity to treat many diseases, or prevent disease progression. Specific targeting sig­
nificantly reduces off-target effects, while simultaneously blocking the pathway that produces
the inflammatory dysregulation, as opposed to a generic block of non-specific events, some of
which are anti-inflammatory. Together, the discovery of JAK inhibitors to block production
of excessive inflammation at the root provides an attractive and still-blossoming field. The

257
258 JAK-STAT Signaling in Diseases

long-term ramifications as they relate to approval for disease indications, the impact of JAK
inhibitors long-term within distinct populations, and the ability of these agents to control
chronic inflammation is still being explored. Despite the relatively new discovery of JAK
inhibitors, and the ongoing research to better understand how these agents can mitigate
disease, multiple approvals of JAK inhibitors have transpired, and several novel JAK
inhibitors are in late-phase clinical investigation. This review focuses on the details of each
agent, the phase of investigation, the target indication, and the mechanistic block of each
JAK as it relates to the downstream disease and inflammatory factors targeted. The dynamic
interaction between these events is discussed, with a goal of broadening the overall knowl­
edge of JAK inhibitors toward treatment, prevention, or eradication of diseases driven by
inflammatory dysregulation.

14.1.1 JAK-STAT Pathway and Downstream Cytokine Production


A wide variety of ligands bind to type 1/2 cytokine receptors, including growth factors/colony­
stimulating factors, interferons, and HIV-1 envelope gp120 (4–10) (Figure 14.1). The resulting
activation of these receptors by bound ligand results in activation of JAKs, comprised of JAK1,
JAK2, JAK3, and Tyk2 (Figure 14.2). JAKs then engage in cross-phosphorylation (one JAK
phosphorylates the other JAK), along with phosphorylation of the intracellular domain of the

FIGURE 14.1 JAK-STAT pathway signaling pathway and downstream activation of proinflammatory cytokines.
JAK-STAT Pathway in Inflammatory Diseases 259

FIGURE 14.2 Summary of JAK inhibitors, specificity for JAKs, and corresponding block on proinflammatory cytokines.

receptor, the latter of which results in docking sites for various recruited signaling molecules
(5–10). Activated phosphorylated JAKs then phosphorylate STATs (Figure 14.2). The STAT
family is comprised of STAT1, STAT2, STAT3, STAT4, STAT5a, STAT5b, and STAT6
(5–10). Activated, dimerized phosphorylated STATs then translocate to the nucleus, and bind
to specific transcription sites within the DNA that confer transcription of gene products for
a wide array pro-inflammatory and immunomodulatory cytokines. Different receptors can
trigger with specific JAKs, thereby agonist engagement at the receptor level can lead to
different downstream cascades of eventual cytokine production. Therefore, based on the
desired blockade of cytokines, selective inhibition of a specific JAK can be sought. None­
theless, there is some overlap in cytokines that are eventually produced from phosphorylation
of JAKs. For example, JAK1 and JAK2 phosphorylation can result in production of IL-6, and
IL-7 is produced after JAK1 or JAK3 phosphorylation (Figure 14.2). A summary of JAK
phosphorylation and downstream production of cytokines appears in Figure 14.2. Notably,
autocrine and paracrine activation by cytokines produced from JAK-STAT triggering can
further activate JAK-STAT signaling, recruitment of bystander cells, and promote a systemic
or tissue-localized inflammatory response (5–10), underscoring the importance of this pathway
in inflammatory disease pathologies.

14.1.2 Diseases Targeted with JAKinibs


Many diseases are hallmarked by chronic inflammation, driven by a dysregulated autoimmune
phenotype, including rheumatoid arthritis, psoriasis, and systemic lupus Erythmeus, Castleman’s
disease (pediatrics), and Chron’s diseases (7–18). Additionally, dysregulated JAK-STAT signaling,
governed by overexpressing JAKs, have been reported in myeloproliferative malignancies including
myelofibrosis and polycythemia vera (9,16–18).
Until recently, dysregulated autoimmune diseases were treated with non-specific steroid-based
regimens or immunosuppressants intended to blunt the inflammatory response. These agents are
blunt tools that systemically reduce the ability of the adaptive and innate immunity to mount
260 JAK-STAT Signaling in Diseases

a response to pathogens, and are also present with off-target toxicities and malignancies
(8,10,12,19,20). A major unmet clinical need is specific blockade of the pro-inflammatory
cytokines that are dysregulated and overexpressed in these conditions. Further, prior to the use
of JAK inhibitors, JAK-overexpressing cancers, such as myelofibrosis and polycythemia vera,
were considered incurable cancers with little more than palliative care as therapeutic interventions.
The discovery of JAK inhibitors to treat inflammation and also block JAK overexpression that
drives cancer has provided the field with a revolutionary modality to tackle previously incurable
diseases with significantly reduced toxicity and an increased quality of life. These discoveries have
lead to exploratory consideration for JAK inhibitors for more than 47 autoimmune and
inflammatory-based disorders (clinicaltrials.gov), including chronic inflammation in HIV-infected
individuals (clinicaltrials.gov/ruxolitinib.a5336; actg.gov/a5336). A closer look at the role of JAK
inhibitors in treating or reversing specific disease pathologies are discussed below.

14.1.3 Arthritis
JAK-STAT dysregulation transpires in arthritis, including rheumatoid arthritis (RA) and psoriatic
arthritis (PsA). PsA is hallmarked by musculoskeletal inflammation, resulting in chronic inflamma­
tion of joints and the systemic musculoskeletal system (10,21–24). Pathogenesis is associated with
multiple factors, including mutations to the human leukocyte antigens (HLA-Cw0602 and HLA­
AB*27 alleles); these mutations are associated with increased dysregulation of inflammatory
cytokines, including tumor necrosis factor alpha (TNF-α, IL-18, IL-17, IL-α/β, and IL-6). TNF-α,
in particular, confers autocrine and paracrine signaling across myeloid and lymphoid cells, promot­
ing production of chemotactic factors that induce localization and recruitment of a Th1 cytokine
response and induction of antigen-presenting cell (APC) inflammatory responses (10,21–24).
Production of these cytokines is triggered by activation of the JAK-STAT pathway; therefore,
blocking its activation with JAK inhibitors represents an attractive and specific modality to prevent
clinical manifestation of symptoms in individuals with PsA. RA is an autoimmune arthritis found
in 1% of the total population, hallmarked by systemic joint inflammation conferred by JAK-STAT
activation and production of IL-6, TNF-α, IL-α/β, D-dimer, C reactive protein (CRP), and other
pro-inflammatory cytokines (21,23,24). JAK-STAT signaling is a primary modulator of production
of these cytokines (Figure 14.2). Prior to the discovery of JAK inhibitors, treatment for RA
centered around blunt steroidal anti-inflammatory regimens, without blocking the root cause of
the signaling dysregulation. More recently, biologics targeting circulating inflammatory cytokines,
such as anti-TNF-α monoclonal antibody (mAb) or anti-IL-1 mAb, remained a more attractive
option versus a systemic and non-specific blunting of systemic immune function. Biologics present
with limitations, however, including lack of oral bioavailability and eventual tolerance from the
immune system, were compounded by lack of multi-cytokine targeting. These limitations led to
pursuit of orally available, targeted inhibitors of multiple pro-inflammatory cytokines by blocking
JAK phosphorylation. Tofacitinib was the first JAK inhibitor approved for this indication, and is
a JAK3 inhibitor that also demonstrates block of other JAKs with lower specificity (summarized in
Figure 14.2). JAK3 blockade has been associated with reduced neutrophil and natural killer cell
counts; therefore, selective JAK1 or 2 blockade without significant JAK3 block has begun to be
considered. To date, baricitinib, a JAK1/2 inhibitor, is also FDA approved for treatment of RA;
baricitinib is a qd-dosed orally bioavailable agent that is also approved for use in children
(summarized in Figure 14.2) (8,10). Multiple investigational JAK inhibitors are also under investi­
gation for rheumatoid arthritis, including peficitinib (phase 3; pan JAK inhibitor), filgotinib (phase
3; JAK1 inhibitor), and upadcitinib (phase 3; JAK1 inhibitor); summarized in Figure 14.2 (8,10).

14.1.4 Inflammatory Bowel Disease


Inflammatory bowel disease (IBD) is delineated into two different disease pathologies: Crohn’s
disease (CD) and ulcerative colitis (UC) (19). Both are comprised of localized inflammation
JAK-STAT Pathway in Inflammatory Diseases 261

within the intestine that results in clinical manifestation of symptoms including diarrhea,
abdominal pain, and fistulas/abscesses (12,19). Polymorphisms in JAK2, Tyk2, and STAT3 genes
have been associated with increased risk of developing IBD, and these JAKs and STATs are
crucial parts of downstream signaling that produces inflammatory cytokines that result in IBD.
Further, the inflammatory responses are conferred in T cells and T cell subsets in response to
JAK-STAT-mediated inflammatory induction compounds’, Th1, Th2, and Th17, dysregulation
and inflammation within the intestine of IBD patients (12,19). Therefore, safe, specific blockade
of JAK-STAT signaling in the context of IBD represents an attractive modality to alleviate
symptoms of this disease. To date, the JAK3 inhibitor, tofacitinib, is approved for this indication,
and multiple JAK inhibitors are under investigation, including filgotinib (phase 3; JAK1 inhibitor
for Crohn’s disease and ulcerative colitis), upadcitinib (phase 2; JAK1 inhibitor for Crohn’s
disease and ulcerative colitis), PF-06651600 (phase 2; JAK3 inhibitor for Crohn’s disease and
ulcerative colitis), and PF-06700841 (JAK1 and Tyk2 inhibitor; phase 2 for Crohn’s disease and
ulcerative colitis) (8,10).

14.1.5 Myeloproliferative Disorders


Polycythemia vera (PV) and myelofibrosis (MF) are two distinct malignancies that are character­
ized by BCR-ABL1-negative myeloproliferations associated with hematopoietic mutations and
hyperactivation of the JAK-STAT pathway (14,16). Dysregulated JAK-STAT signaling disrupts
hematopoietic growth factors, including proliferation of stem cells and production of erythro­
poietin/thrombopoietin (14,16); this is hallmarked in both MF and PV. Further, a gain of function
mutation in the JAK2 gene (JAK2V617F) promotes constitutive activation of JAK2, resulting in
elevated levels of inflammatory cytokines, including IL-6, TNF-α, IL-1, CRP, and D-dimer
(14,16). Presence of abnormally high levels of these cytokines is thought to further promote
autocrine and paracrine stimulation to produce these cytokines in non-JAK2 mutated cells,
further compounding the hyperstimulatory profile of MF and PV. This chronic triggering of pro-
inflammatory cytokines is thought to drive the clinical manifestation of symptoms in patients with
MF and PV, further underscoring necessity of JAK-STAT blockade in treatment of myeloproli­
ferative disorders. To date, ruxolitinib, a JAK1/2 inhibitor, is the only FDA-approved JAK
inhibitor for treatment of PV and MF (16,18,25), but multiple other JAK inhibitors are under
various phases of clinical investigation for myeloproliferative disorders, including momelotinib
(phase 3; JAK2 inhibitor), fedratinib (phase 2; JAK2 inhibitor), and gandotinib (phase 2; JAK2
inhibitor), summarized in Figure 14.2 (8,10).

14.1.6 Psoriasis and Autoimmune Skin Diseases


Inflammatory skin diseases are hallmarked by cutaneous pro-inflammatory milieu in localized
APCs (dendritic cells, macrophages, etc.), and cells of the adaptive immune response (Th cells and
B cells) (20). Localization of these cells to the skin provides a robust innate and adaptive immune
response, but can become dysregulated in chronic inflammatory skin diseases, often by autoim­
mune-based mechanisms. This results in elevated pro-inflammatory cytokines, localized within the
dermis, including elevated levels of IL-17, and IL-22 (psoriasis) or IL-4 (atopic dermatitis) (20).
Most inflammatory skin diseases, independent of origin, were treated up until recent with topical
steroids or oral systemic immunosuppressive agents designed to blunt the chronic inflammation.
These treatments do not treat the root signaling cascades that result in production of inflammatory
cytokines, but instead only treat the phenotype of inflamed skin. Recent advances have defined the
critical role of the JAK-STAT pathway in (1) producing the inflammatory milieu found in
inflammatory skin disorders (8,20), and (2) dysregulation and genetic alterations found in auto­
immune skin disorders (8,10). Together, these findings have resulted in exploration of JAK
inhibitors for treatment of inflammatory skin disorders, including psoriasis. To date, tofacitinib
(JAK3 inhibitor with some specificity for other JAKs) is FDA approved for psoriatic arthritis (21),
262 JAK-STAT Signaling in Diseases

and multiple other JAK inhibitors are under various phases of clinical investigation for inflammatory
skin disorders, including ruxolitinib (phase 2; JAK1/2 inhibitor for atopic dermatitis), baricitinib
(JAK1/2 inhibitor for atopic dermatitis), PF-06263276 (pan JAK inhibitor; phase 1 for psoriasis),
upadcitinib (JAK1 inhibitor; phase 2 atopic dermatitis), and PF-04965842 (JAK1 inhibitor; phase 2
for psoriasis, phase 3 for atopic dermatitis), and BMS- 986165 (Tyk2 inhibitor; phase 2 for psoriasis),
summarized in Figure 14.2 (8,10).

14.1.7 Systemic Lupus Erythematosus (SLE)


SLE is a systemic autoimmune disorder affecting ~5 million people worldwide, which is char­
acterized by high levels of circulating interferons that produce systemic inflammatory immune
dysregulation. Together, this can result in massive autoimmune inflammatory responses against
major tissues and organs, including skin and vital organs; left unchecked during flare-ups, this
disease can lead to death (8,10,20). Treatments for this condition previously included systemic
immunosuppressive agents, including cyclosporins and methotrexate, each of which is present
with significant toxicities and an overall immunosuppressive phenotype due to lack of specificity
for the disease pathology. A major unmet clinical need for SLE is safe, specific potent block of
pro-inflammatory cytokine production, making JAK inhibitors, in particular JAK1/2 inhibitors
that block the IFN response, very attractive. To that end, in December 2018, it was announced
that baricitinib, an orally bioavailable qd-dosed JAK1/2 inhibitor (currently FDA-approved for
rheumatoid arthritis), received fast-track status from the FDA for treatment of SLE (fda.gov;
elililly.com). Data from this trial will inform the field about the impact of JAK1/2 inhibition in
downregulating the chronic inflammation hallmarking SLE, and provide a foundation for further
studies toward careful evaluation of JAK inhibitors for this indication.

14.1.8 Alopecia and Hair Loss


Alopecia areata (AA) is an autoimmune disease with no currently FDA-approved therapies; the
condition is characterized by non-scarring hair loss (26). The full mechanism(s) responsible for this
phenotype is/are not well understood; however, recent discoveries elucidated over active JAK-STAT
signaling in patients with AA (26). Recently, a multi-center study (ClinicalTrials.gov NCT02197455
and NCT02312882) evaluated the ability of tofacitinib, a pan JAK inhibitor with JAK3 specificity,
to reverse AA (27,28). Data demonstrated that tofacitinib was well-tolerated, safe, and effective for
patients with severe AA; however, the treatment length did not demonstrate a durable and sustained
reversal of AA (27,28). Nonetheless, this important study demonstrated encouraging results, and
provided a foundation for future studies with other JAK inhibitors, including JAK1/2 inhibitors, to
evaluate their ability to reverse AA. In particular, a recent report was described for a single patient
entering a trial for baricitinib for treatment of CANDLE (chronic atypical neutrophilic dermatitis
with lipodystrophy and elevated temperature). The patient who entered the study also had AA; the
report describes a reversal of AA in this individual (29). Although this report is a sample size of
one, it underscores the potential for JAK inhibitors toward reversal of AA or hair loss, especially in
light of the key role of JAK-STAT signaling in conferring the disease pathology in vivo.

14.1.9 Closer Look at JAKinibs Toward HIV-1 Eradication


Despite major advances in treatment of HIV infection, including sustained viral suppression with
highly active antiretroviral therapy (HAART), purge of the HIV reservoir remains challenging.
A major barrier to eradication of HIV is the maintenance of the HIV reservoir (30–33). Recent
reports demonstrate that elevated levels of inflammatory and immunomodulatory cytokines that
trigger through the JAK-STAT pathway, including IL-7 and IL-15 promote proliferation of the
HIV reservoir (homeostatic proliferation) (31,34). Further, other inflammatory cytokines that are
produced from the JAK-STAT pathway remain elevated in HIV-infected individuals, even with
JAK-STAT Pathway in Inflammatory Diseases 263

well-controlled viremia, and are associated with non-AIDS-related morbidity/mortality, HIV


persistence, progression to AIDS, and death (13,35–40). Recent reports demonstrate that JAK
inhibitors, tofacitinib (JAK3, some specificity for JAK1 and 2) and ruxolitinib (JAK1/2), block
HIV reservoir establishment, maintenance, and expansion in vitro and ex vivo (41). Further, these
inhibitors downregulate key markers of HIV persistence ex vivo, underscoring the potential of
JAK inhibitors as an add-on therapy to block key events that prevent HIV eradication in vivo,
which could result in an eventual purge of the HIV reservoir in vivo (Figure 14.3A) (41). The
impact of ruxolitinib on inflammation associated with HIV infection, which plays an important
role in seeding viral reservoirs, viral persistence, and ongoing low-level replication in sanctuary
sites is currently being evaluated in an NIH-ACTG sponsored 21-site Phase 2a study in humans
(A5336; https://actgnetwork.org/newsletter/april2016/A5336); data from this study will be
reported at the CROI meeting in 2019. Briefly, the population studied are “otherwise healthy”
HIV-infected individuals, with CD4 T cell counts >350 mm3, no underlying disease or co-
infections, and no elevated inflammatory markers at onset of study. This population was selected
because of the “proof-of-concept” nature of A5336; the goal of the study was to evaluate safety of
an immunomodulator-based agent in the context of HIV infection, and further to quantify any
potential decreases in inflammatory or HIV-persistence markers within the population. It is well
established that chronic inflammation is associated with end-organ disease and mortality for
people living with HIV (PLWH); and ruxolitinib demonstrates reduction in biomarkers of
systemic inflammation in HIV-uninfected individuals and HIV reservoir and persistence markers
ex vivo. The goal of this trial was to determine the safety and efficacy of ruxolitinib in treated
HIV disease. In a highly selected cohort of HIV-positive adults on suppressive ART, ruxolitinib
was safe and well tolerated, and demonstrated a trend in reduction of IL-6 levels within the short
duration of treatment (5-week intervention). These data were coupled with a statistically significant

FIGURE 14.3 Role of the JAK-STAT pathway in HIV-1 persistence, HIV-associated neurocognitive dysfunction, and
HIV-associated cardiovascular disease. Elevated pSTATS correlates with increased HIV reservoirs, HIV persistence
markers, and progression to AIDS in HIVinfected individuals (A). Increased JAK-STAT activation is associated with
HIV-associated neurocognitive dysfunction (HAND), which is hallmarked by CNS inflammation that promotes
HAND (B). Increased inflammation in myeloid cells in HIV-infected individuals results in increased atherosclerosis;
key markers including sCD163 are associated with increased risk of CVD in HIV-infected individuals (C).
264 JAK-STAT Signaling in Diseases

decrease in sCD14 levels, and an increase in circulating T cells through mechanisms undefined. This
proof-of-concept trial provides a rationale for future studies of JAK inhibitors in PLWH, who have
residual inflammation or immune dysfunction despite long-term suppressive ART, including in
individuals with HIV-associated neurocognitive dysfunction, and HIV-associated cardiovascular
abnormalities and atherosclerosis.
Dysregulation of the JAK-STAT pathway has also been associated with HAND (6,15,38,42–
45), and ruxolitinib demonstrated amelioration of HIV-induced inflammation in a murine model
(46). Further, recent reports demonstrate that baricitinib crosses the blood–brain barrier and
reverses behavioral abnormalities associated with HAND in a murine model, and reverses key
pathological markers of HAND (47) (Figure 14.3B). Given these findings, JAK inhibitors,
especially baricitinib, which demonstrates a safe, qd dosing in humans with FDA approval in
children, could be used with HAART to reverse HAND and key events driving HIV persistence;
further studies in humans are warranted for this indication.
Increased inflammation that persists in HIV-infected individuals on otherwise effective therapy
have a higher-than-expected risk of a number of “non-AIDS” conditions, including cardiovascular
disease (6,35,36,38–40,42,48–50). Chronic inflammation predicts and presumably causes this
effect, and elevated activation of the JAK-STAT pathway, which drives a pro-inflammatory
milieu, has been reported during HIV infection and is associated with HIV persistence and
cardiovascular disease in HIV-infected individuals (summarized in Figure 14.3) (6,35,36,38–
40,42,48–50). Mitigating increased risk of CVD in HIV-infected individuals, which data suggests
is driven by increased, persistent elevated levels of inflammation, is an unmet clinical need. Safe,
specific potent inhibition of these events with a well-tolerated, safe add-on therapy could reduce
the risk of CVD in HIV-infected individuals. Given the body of data regarding the potent, specific
block that JAK inhibitors confer on inflammatory cytokines, and their anti-HIV properties, it is
possible that use of JAK inhibitors to reduce CVD in HIV-infected individuals could represent an
attractive modality for a currently unaddressed clinical indication.

14.1.10 Metabolism and Future Indications


Metabolism of JAK inhibitors relative to their clearance and potential drug–drug interactions is an
important factor for consideration of these agents when co-administered with other agents. Agents
that are orally administered have two major routes of metabolism; renal or hepatic (51,52). Hepatic
clearance often presents with drug–drug interactions, due to potential competition for key metabo­
lizing enzymes such as CYP3A4 and CYP2B6; occupation of the CYP-binding sites for metabolism
by a JAK inhibitor can prevent another drug from binding, and therefore prevent its metabolism
(51,52). The end result of such an interaction is a potentially toxic concentration of an agent, when
co-administered in the presence of a JAK inhibitor. Ruxolitinib is a hepatic cleared JAK inhibitor,
for example (18,53,54); therefore, careful consideration for co-administration of agents that also
require hepatic metabolism and clearance by CYP3A4 and CYP2B6 must be carefully monitored, or
avoided altogether. Second generation JAK inhibitors, such as baricitinib (Olumiant, Eli Lilly), are
renally cleared, which avoid the potential drug–drug interaction at the level of CYP metabolism
(53,54), therefore significantly limiting the list of agents that may have drug–drug interactions with
baricitinib versus ruxolitinib (or another hepatically cleared JAK inhibitor). Further, recent reports
have begun to evaluate the pharmacokinetics of JAK inhibitors across different compartments,
including the CNS (46,47), providing the beginning of exploration of use of these agents for tissue or
compartment-specific pathologies, including neuroinflammatory diseases.
An additional indication that is not currently being explored clinically is atherosclerosis, which is
driven by a pro-inflammatory milieu within cardiac tissue, including sCD163, sCD14, and localized
inflammation from resident myeloid cells within the heart (39). Decreasing persistent tissue-specific
inflammation for this indication could provide an add-on therapy to reverse or prevent cardiac
events that lead to morbidity and mortality in the aging population. Atherosclerosis is increased in
the aging HIV-positive population, and key markers of soluble inflammation have been correlated
JAK-STAT Pathway in Inflammatory Diseases 265

with increased morbidity and mortality from non-AIDS-related events including heart attacks (39),
further highlighting the potential for this indication across the ~39 million HIV-infected individuals
worldwide.

14.1.11 Summary
The discovery of safe, JAK–specific inhibitors that are orally bioavailable and well-tolerated has
provided a new chapter toward harnessing of the JAK-STAT pathway in many diseases. Myelopro­
liferative disorders previously had no FDA-approved therapies prior to JAK inhibitors, resulting in
rapid progression of disease and death. Other conditions including many autoimmune disorders
hallmarked by chronic uncontrolled inflammation were treated with blunt tools, such as systemic
immunosuppressants, which resulted in significant toxicities. The onset of JAK inhibitors to treat
inflammatory conditions marks the beginning of a burgeoning new field. Exploration of the effect of
safe, specific inhibition of JAKs toward immunomodulation of a wide array of inflammatory-driven
diseases, including viral infections, will continue to be elucidated. New information gained from
clinical trials will guide eventual additional FDA approvals of novel, well-tolerated, and effective
JAK inhibitors.

ACKNOWLEDGMENTS
This work was funded in part by NIH grant 1RO1-MH-116695 and Emory’s CFAR NIH grant
P30-AI-050409.

REFERENCES
1. Lei J, Zhou Y, Xie D, Zhang Y. 2015. Mechanistic insights into a classic wonder drug–aspirin. J Am
Chem Soc 137:70–73.
2. Ornelas A, Zacharias-Millward N, Menter DG, Davis JS, Lichtenberger L, Hawke D, Hawk E,
Vilar E, Bhattacharya P, Millward S. 2017. Beyond COX-1: the effects of aspirin on platelet biology
and potential mechanisms of chemoprevention. Cancer Metastasis Rev 36:289–303.
3. Shim YK, Kim N. 2016. [Nonsteroidal anti-inflammatory drug and aspirin-induced peptic ulcer
disease]. Korean J Gastroenterol 67:300–312.
4. Kohler JJ, Tuttle DL, Coberley CR, Sleasman JW, Goodenow MM. 2003. Human immunodeficiency
virus type 1 (HIV-1) induces activation of multiple STATs in CD4+ cells of lymphocyte or monocyte/
macrophage lineages. J Leukoc Biol 73:407–416.
5. Benczik M, Gaffen SL. 2004. The interleukin (IL)-2 family cytokines: survival and proliferation
signaling pathways in T lymphocytes. Immunol Invest 33:109–142.
6. Chaudhuri A, Yang B, Gendelman HE, Persidsky Y, Kanmogne GD. 2008. STAT1 signaling
modulates HIV-1-induced inflammatory responses and leukocyte transmigration across the
blood-brain barrier. Blood 111:2062–2072.
7. Darnell JE, Jr., Kerr IM, Stark GR. 1994. JAKJAK–STAT pathways and transcriptional activation
in response to IFNs and other extracellular signaling proteins. Science 264:1415–1421.
8. Gadina M, Johnson C, Schwartz D, Bonelli M, Hasni S, Kanno Y, Changelian P, Laurence A,
O’Shea JJ. 2018. Translational and clinical advances in JAK-STAT biology: the present and future of
jakinibs. J Leukoc Biol 104:499–514.
9. McLornan DP, Khan AA, Harrison CN. 2015. Immunological consequences of JAK inhibition:
friend or foe? Curr Hematol Malig Rep 10:370–379.
10. O’Shea JJ, Schwartz DM, Villarino AV, Gadina M, McInnes IB, Laurence A. 2015. The JAK-STAT
pathway: impact on human disease and therapeutic intervention. Annu Rev Med 66:311–328.
11. Bovolenta C, Camorali L, Lorini AL, Ghezzi S, Vicenzi E, Lazzarin A, Poli G. 1999. Constitutive
activation of STATs upon in vivo human immunodeficiency virus infection. Blood 94:4202–4209.
12. Coskun M, Salem M, Pedersen J, Nielsen OH. 2013. Involvement of JAK/STAT signaling in the
pathogenesis of inflammatory bowel disease. Pharmacol Res 76:1–8.
266 JAK-STAT Signaling in Diseases

13. Gavegnano C, Detorio M, Montero C, Bosque A, Planelles V, Schinazi RF. 2014. Ruxolitinib and
tofacitinib are potent and selective inhibitors of HIV-1 replication and virus reactivation in vitro.
Antimicrob Agents Chemother 58:1977–1986.
14. Griesshammer M, Sadjadian P. 2017. The BCR-ABL1-negative myeloproliferative neoplasms: a review
of JAK inhibitors in the therapeutic armamentarium. Expert Opin Pharmacother 18:1929–1938.
15. Nabavi SM, Ahmed T, Nawaz M, Devi KP, Balan DJ, Pittala V, Arguelles-Castilla S, Testai L,
Khan H, Sureda A, de Oliveira MR, Vacca RA, Xu S, Yousefi B, Curti V, Daglia M, Sobarzo-
Sanchez E, Filosa R, Nabavi SF, Majidinia M, Dehpour AR, Shirooie S. 2018. Targeting STATs in
neuroinflammation: the road less traveled! Pharmacol Res 141:73–84.
16. Bryan JC, Verstovsek S. 2016. Overcoming treatment challenges in myelofibrosis and polycythemia
vera: the role of ruxolitinib. Cancer Chemother Pharmacol 77:1125–1142.
17. Massa M, Rosti V, Campanelli R, Fois G, Barosi G. 2014. Rapid and long-lasting decrease of
T-regulatory cells in patients with myelofibrosis treated with ruxolitinib. Leukemia 28:449–451.
18. Verstovsek S, Mesa RA, Gotlib J, Levy RS, Gupta V, DiPersio JF, Catalano JV, Deininger M, Miller C,
Silver RT, Talpaz M, Winton EF, Harvey JH, Jr., Arcasoy MO, Hexner E, Lyons RM, Paquette R,
Raza A, Vaddi K, Erickson-Viitanen S, Koumenis IL, Sun W, Sandor V, Kantarjian HM. 2015. A
double-blind, placebo-controlled trial of ruxolitinib for myelofibrosis. N Engl J Med 366:799–807.
19. Chan HC, Ng SC. 2017. Emerging biologics in inflammatory bowel disease. J Gastroenterol
52:141–150.
20. Shreberk-Hassidim R, Ramot Y, Zlotogorski A. 2017. Janus kinase inhibitors in dermatology:
a systematic review. J Am Acad Dermatol 76:745–753e719.
21. Costa L, Del Puente A, Peluso R, Tasso M, Caso P, Chimenti MS, Sabbatino V, Girolimetto N,
Benigno C, Bertolini N, Del Puente A, Perricone R, Scarpa R, Caso F. 2017. Small molecule therapy
for managing moderate to severe psoriatic arthritis. Expert Opin Pharmacother 18:1557–1567.
22. Song GG, Bae SC, Lee YH. 2014. Efficacy and safety of tofacitinib for active rheumatoid arthritis
with an inadequate response to methotrexate or disease-modifying antirheumatic drugs: a
meta-analysis of randomized controlled trials. Korean J Intern Med 29:656–663.
23. Tanaka Y. 2016. [Tofacitinib for the treatment of rheumatoid arthritis]. Nihon Rinsho 74:974–980.
24. Winthrop KL. 2017. The emerging safety profile of JAK inhibitors in rheumatic disease. Nat Rev
Rheumatol 13:234–243.
25. Chen X, Williams WV, Sandor V, Yeleswaram S. 2013. Population pharmacokinetic analysis of
orally-administered ruxolitinib (INCB018424 Phosphate) in patients with primary myelofibrosis
(PMF), post-polycythemia vera myelofibrosis (PPV-MF) or post-essential thrombocythemia myelo­
fibrosis (PET MF). J Clin Pharmacol 53:721–730.
26. Xing L, Dai Z, Jabbari A, Cerise JE, Higgins CA, Gong W, de Jong A, Harel S, DeStefano GM,
Rothman L, Singh P, Petukhova L, Mackay-Wiggan J, Christiano AM, Clynes R. 2014. Alopecia
areata is driven by cytotoxic T lymphocytes and is reversed by JAK inhibition. Nat Med 20:1043–1049.
27. Kennedy Crispin M, Ko JM, Craiglow BG, Li S, Shankar G, Urban JR, Chen JC, Cerise JE,
Jabbari A, Winge MC, Marinkovich MP, Christiano AM, Oro AE, King BA. 2016. Safety and
efficacy of the JAK inhibitor tofacitinib citrate in patients with alopecia areata. JCI Insight 1:e89776.
28. Samadi A, Ahmad Nasrollahi S, Hashemi A, Nassiri Kashani M, Firooz A. 2017. Janus kinase
(JAK) inhibitors for the treatment of skin and hair disorders: a review of literature. J Dermatolog
Treat 28:476–483.
29. Jabbari A, Dai Z, Xing L, Cerise JE, Ramot Y, Berkun Y, Sanchez GA, Goldbach-Mansky R,
Christiano AM, Clynes R, Zlotogorski A. 2015. Reversal of alopecia areata following treatment with
the JAK1/2 inhibitor baricitinib. EBioMedicine 2:351–355.
30. Barouch DH, Deeks SG. 2014. Immunologic strategies for HIV-1 remission and eradication. Science
345:169–174.
31. Chomont N, El-Far M, Ancuta P, Trautmann L, Procopio FA, Yassine-Diab B, Boucher G,
Boulassel MR, Ghattas G, Brenchley JM, Schacker TW, Hill BJ, Douek DC, Routy JP,
Haddad EK, Sekaly RP. 2009. HIV reservoir size and persistence are driven by T cell survival and
homeostatic proliferation. Nat Med 15:893–900.
32. Gavegnano C, Schinazi RF. 2009. Antiretroviral therapy in macrophages: implication for HIV
eradication. Antivir Chem Chemother 20:63–78.
JAK-STAT Pathway in Inflammatory Diseases 267

33. Honeycutt JB, Thayer WO, Baker CE, Ribeiro RM, Lada SM, Cao Y, Cleary RA, Hudgens MG,
Richman DD, Garcia JV. 2017. HIV persistence in tissue macrophages of humanized myeloid-only
mice during antiretroviral therapy. Nat Med 23:638–643.
34. Vandergeeten C, Fromentin R, DaFonseca S, Lawani MB, Sereti I, Lederman MM, Ramgopal M,
Routy JP, Sekaly RP, Chomont N. 2013. Interleukin-7 promotes HIV persistence during antiretro­
viral therapy. Blood 121:4321–4329.
35. Bastard JP, Soulie C, Fellahi S, Haim-Boukobza S, Simon A, Katlama C, Calvez V, Marcelin AG,
Capeau J. 2016. Circulating interleukin-6 levels correlate with residual HIV viraemia and markers of
immune dysfunction in treatment-controlled HIV-infected patients. Antivir Ther 17:915–919.
36. Borges AH, O’Connor JL, Phillips AN, Baker JV, Vjecha MJ, Losso MH, Klinker H, Lopardo G,
Williams I, Lundgren JD, Group ISS, Group ES, Committee SS. 2014. Factors associated with
D-dimer levels in HIV-infected individuals. PLoS One 9:e90978.
37. Chakrabarti LA, Boucherie C, Bugault F, Cumont MC, Roussillon C, Breton G, Patey O, Chene G,
Richert L, Lortholary O, Anrs 129 Bkvir-Cytok Study Group. 2014. Biomarkers of CD4+ T-cell
activation as risk factors for tuberculosis-associated immune reconstitution inflammatory syndrome.
AIDS 28:1593–1602.
38. Chaudhuri A, Duan F, Morsey B, Persidsky Y, Kanmogne GD. 2008. HIV-1 activates proinflamma­
tory and interferon-inducible genes in human brain microvascular endothelial cells: putative
mechanisms of blood-brain barrier dysfunction. J Cereb Blood Flow Metab 28:697–711.
39. Frustaci A, Petrosillo N, Vizza D, Francone M, Badagliacca R, Verardo R, Fedele F, Ippolito G,
Chimenti C. 2014. Myocardial and microvascular inflammation/infection in patients with
HIV-associated pulmonary artery hypertension. AIDS 28:2541–2549.
40. Tenorio AR, Zheng Y, Bosch RJ, Krishnan S, Rodriguez B, Hunt PW, Plants J, Seth A, Wilson CC,
Deeks SG, Lederman MM, Landay AL. 2014. Soluble markers of inflammation and coagulation but
not T-cell activation predict non-AIDS-defining morbid events during suppressive antiretroviral
treatment. J Infect Dis 210:1248–1259.
41. Gavegnano C, Brehm JH, Dupuy FP, Talla A, Ribeiro SP, Kulpa DA, Cameron C, Santos S,
Hurwitz SJ, Marconi VC, Routy JP, Sabbagh L, Schinazi RF, Sekaly RP. 2017. Novel mechanisms
to inhibit HIV reservoir seeding using JAK inhibitors. PLoS Pathog 13:e1006740.
42. Clayton KL, Garcia JV, Clements JE, Walker BD. 2017. HIV infection of macrophages: implications
for pathogenesis and cure. Pathog Immun 2:179–192.
43. Dunfee R, Thomas ER, Gorry PR, Wang J, Ancuta P, Gabuzda D. 2006. Mechanisms of HIV-1
neurotropism. Curr HIV Res 4:267–278.
44. Sippy BD, Hofman FM, Wallach D, Hinton DR. 1995. Increased expression of tumor necrosis
factor-alpha receptors in the brains of patients with AIDS. J Acquir Immune Defic Syndr Hum
Retrovirol 10:511–521.
45. Tyor WR, Glass JD, Griffin JW, Becker PS, McArthur JC, Bezman L, Griffin DE. 1992. Cytokine
expression in the brain during the acquired immunodeficiency syndrome. Ann Neurol 31:349–360.
46. Haile WB, Gavegnano C, Tao S, Jiang Y, Schinazi RF, Tyor WR. 2016. The Janus kinase inhibitor
ruxolitinib reduces HIV replication in human macrophages and ameliorates HIV encephalitis in
a murine model. Neurobiol Dis doi:10.1016/j.nbd.2016.02.007.
47. Gavegnano C, Haile W, Koneru R, Hurwitz S, Tao S, Tyor WR, Schinazi RF. 2018. Baricitinib
reverses HAND behavioral phenotype in vivo and reservoir seeding in vitro. Keystone Symposium,
Whistler, BC, April 2018.
48. Buzon MJ, Massanella M, Llibre JM, Esteve A, Dahl V, Puertas MC, Gatell JM, Domingo P,
Paredes R, Sharkey M, Palmer S, Stevenson M, Clotet B, Blanco J, Martinez-Picado J. 2010. HIV-1
replication and immune dynamics are affected by raltegravir intensification of HAART-suppressed
subjects. Nat Med 16:460–465.
49. Kamat A, Misra V, Cassol E, Ancuta P, Yan Z, Li C, Morgello S, Gabuzda D. 2015. A plasma
biomarker signature of immune activation in HIV patients on antiretroviral therapy. PLoS One 7:
e30881.
50. Katlama C, Deeks SG, Autran B, Martinez-Picado J, van Lunzen J, Rouzioux C, Miller M, Vella S,
Schmitz JE, Ahlers J, Richman DD, Sekaly RP. 2013. Barriers to a cure for HIV: new ways to target
and eradicate HIV-1 reservoirs. Lancet 381:2109–2117.
268 JAK-STAT Signaling in Diseases

51. Knights KM, Rowland A, Miners JO. 2013. Renal drug metabolism in humans: the potential for
drug-endobiotic interactions involving cytochrome P450 (CYP) and UDP-glucuronosyltransferase
(UGT). Br J Clin Pharmacol 76:587–602.
52. Konig J, Muller F, Fromm MF. 2013. Transporters and drug-drug interactions: important determi­
nants of drug disposition and effects. Pharmacol Rev 65:944–966.
53. Shi JG, Chen X, Emm T, Scherle PA, McGee RF, Lo Y, Landman RR, McKeever EG, Jr.,
Punwani NG, Williams WV, Yeleswaram S. 2014. The effect of CYP3A4 inhibition or induction on
the pharmacokinetics and pharmacodynamics of orally administered ruxolitinib (INCB018424
phosphate) in healthy volunteers. J Clin Pharmacol 52:809–818.
54. Shi JG, Chen X, McGee RF, Landman RR, Emm T, Lo Y, Scherle PA, Punwani NG, Williams WV,
Yeleswaram S. 2014. The pharmacokinetics, pharmacodynamics, and safety of orally dosed
INCB018424 phosphate in healthy volunteers. J Clin Pharmacol 51:1644–1654.
Index

α-syn, 88–90 AS1810722, 24


β-catenin, 40 Ascaris, 56
γc, 124, 160, 246 Asthma, 7, 24, 39, 49–59, 247, 248
Astrocyte, 83, 85–87, 90, 92, 93, 95
A ASXL1, 192
Atherosclerosis, 44, 103, 108, 109, 167, 263, 264
A53T mutant, 88 Atopic dermatitis (See AD)
AC-4–130, 23 Atopy, 56, 57, 73, 247
Acanthosis, 70, 75 Autocrine, 85, 92, 129, 196, 197, 199, 225, 229, 259–261
Acetylation, 111, 165, 242 Autoimmune cytopenia, 247
Activation loop, 12, 13, 17–19, 183 Autoimmunity, 52, 162, 166, 243, 246, 247
Acute lymphoblastic leukemia (See ALL) Auto-inhibitory, 11, 13, 190
Acute megakaryoblastic leukemia (See AMKL) Autophagy, 43, 107, 110
Acute myeloid leukemia (See AML) Autophosphorylation, 1, 3, 12, 13, 23, 104, 106,
AD, 69–79 112, 113
Adipocyte, 125–129, 131, 136 Axon, 37, 43, 83
Adipokine, 126, 134 AZD1480, 89, 90, 185, 199
Adiponectin, 109 AZD9150, 232
Adult, 69, 77, 123, 126, 129, 177–179, 181–199, 229, 263 Aβ plaque, 87
Affinity, 6, 14, 19, 38, 105, 114, 168, 242
AG490, 105, 113, 114, 230 B
Agomir, 43
AIDS, 263–265 B cells, 55, 72, 73, 84, 91, 162, 167,
Airway hyperresponsiveness, 50, 51, 53–57 185, 228, 261
Akt, 40, 45, 109, 167, 183, 189, 191, 229 BACE1, 87
ALCL, 194–196 Ba/F3 cell, 181, 183, 184, 189–191, 196
ALK, 194–196 BALB/c, 38, 53, 56
ALL, 7, 177–186, 194, 227, 228, 245 B-ALL, 7, 178–180, 182, 183, 228, 245
Allele, 148, 162, 190, 198, 231, 260 B-ALL, 7, 178–180, 182, 183, 228, 245
Allergic rhinitis, 52, 55, 56, 70 Baricitinib, 76, 77, 160, 162, 163, 165, 248, 260,
Alpha-synuclein, 88, 89 262, 264
Alternaria, 55, 245 Basophils, 74
Alzheimer’s disease, 84, 87, 88 Batf, 242
AMKL, 181, 184 Bax, 21
AML, 7, 23, 178, 179, 183, 184, 186, 225, 227, 228, BBB, 91
230–232 BBI608, 22
Anaplastic large-cell lymphoma (See ALCL) B-cell lymphoma (See BCL)
Anaplastic lymphoma kinase (See ALK) BcL-2, 43
Anemia, 107, 163, 169, 228, 230–232, 248 BCR-ABL, 179, 186, 225, 226, 261
Angiogenesis, 21, 43, 109, 115, 232 Beclin-1, 43
Angiotensin, 112, 114, 164 Berbamine, 23
Angiotensin II, 104, 105, 107, 108, 111–115, 159, 164 Biopsy, 168, 169, 245
Antigen presenting cell, 257, 260 Blood-brain barrier (See BBB)
Anti-inflammatory, 3, 18, 57, 58, 70, 84, 85, 92, 94, 95, BMP, 40
124, 147, 150, 151, 193, 257, 260 BMS- 986165, 262
AP-1, 87 Bone marrow, 39, 40, 92, 107, 167, 178, 189, 191,
Apocrine gland, 70 227–229
Apolipoprotein E, 21 Bortezomib, 229
Apoptosis, 1, 8, 9, 20–23, 38, 40, 43, 87, 107, 109, 110, Boswellic acid, 56
144, 146, 149, 151, 184, 197, 232, 246 Box1, 1
Arginine, 14, 242 Box2, 1
ARIP3 (See PIASx) BP-1-102, 21
AS15174999, 58 BRD0476, 20, 21

269
270 Index

Brentuximab vedotin, 200 Chronic kidney disease (See CKD)


Bromodomain, 185 Chronic lymphocytic leukemia (See CLL)
Brown adipose tissue, 126 Chronic myeloid leukemia (See CML)
Chronic myelomonocytic leukemia (See CMML)
C Chronic Obstructive Pulmonary Disorder (See COPD)
CHZ868, 185, 231
C188-9, 21, 22 Ciliary neurotrophic factor, 4, 7, 84, 135
C1q, 87, 88 CIS, 6, 19, 20, 243
C28/I2, 40 Cisplatin, 21, 22
Calcineurin, 77, 168 cis-regulatory, 242
Calcineurin inhibitor (See CNI) CJ-1383, 22
Calcitonin-1, 7 CKD, 159
CALR, 7, 188, 191, 192, 226, 227 CLCF1, 4, 85, 109
Calreticulin (See CALR) CLL, 229, 231
cAMP, 40 CML, 7, 186, 225, 231
CAN, 168, 169 CMML, 228, 231
Candida albicans, 246 CNI, 168, 169
Carcinogenesis, 149, 150 CNS, 84, 85, 91–93, 129, 134, 135, 263, 264
Cardiac hypertrophy, 36, 39, 108, 110, 111 Codon, 187, 226, 246
Cardiomyocyte, 39, 105, 109–111 Coiled-coil domain, 2, 11, 16, 51
Cardiomyopathy, 106, 110 Collagen, 111, 144, 147, 148, 150, 151, 163, 164
Cardiotrophin-1, 4, 84 COMFORT-I, 230
Cardiotrophin-like cytokine factor 1 (See CLCF1) COMFORT-II, 230
Cartilage, 36, 40 Con A, 18, 144, 145, 147, 148, 150
Caspase-3, 43 Concanavalin A (See Con A)
Caspase-8, 43 COPD, 52, 54
Castleman’s disease, 259 Coronary artery, 36
Caveolin-3, 109 COS-7 cells, 112
CCL11, 53, 57 COX, 257
CCL17, 53, 55, 74 COX-1, 257
CCL2, 85, 93, 109 COX-2, 87, 110, 257
CCL24, 57 CpG island, 38
CCL26, 57 C-reactive protein (See CRP)
CCL5, 53, 94 CRISPR-Cas9, 249
CCL6, 53 CRLF2, 180–182, 185
CD36, 133, 134 Crohn’s disease, 260, 261
CD4, 21 CRP, 260, 261
CD40, 39, 55 CSF1R, 231
CD45, 6, 243 c-SRC, 112
CD8, 107, 247 CUEDC2, 37, 42
CD83, 39 Cullin, 20, 37
CD86, 39 Cullin-2, 6
CDK5R1, 37, 43 Cullin-5, 6, 20
Cell viability, 21, 23, 38, 40, 132, 196, 197, 229 Curcumin, 22, 56
Central nervous system (See CNS) Cutaneous T-cell lymphoma, 194, 195
Cerdulatinib, 195, 199 CXCL1, 85
Cerebellum, 90, 93 CXCL10, 85
Cerebral cortex, 83, 93 CXCL2, 85
c-FLIPL, 110 Cyclooxygenase (See COX)
c-FLIPS, 110 Cyclosporine, 168
c-Fos, 39, 242 CYP2B6, 264
Chemokine, 52, 53, 56, 74, 91, 93, 109, 115, 159, 164, CYP3A4, 24, 264
169, 229 Cytokine, 1–8, 9, 11, 12, 16–20, 37–40, 43–45, 49–53,
Chemotherapy, 22, 178, 179, 185, 199, 232 55–57, 59, 69, 72–75, 78, 79, 84, 85, 87–95,
Cholangiocyte, 149 103–109, 111–113, 115, 124, 126, 129, 134,
Chondrogenic, 40 135, 143–151, 159–169, 177–183, 186, 189,
Chromosome, 36, 73, 93, 149, 179, 182, 191 190, 195–199, 225, 227–229, 232, 233,
Chronic allograft nephropathy (See CAN) 241–248, 257–262, 264
Chronic atypical neutrophilic dermatitis with Cytokine receptor, 1–4, 11, 17, 19, 49, 51, 112, 113, 144,
lipodystrophy and elevated temperature, 262 160, 177, 180, 233, 241, 246, 258
Index 271

Cytokine-inducible SH2 protein (See CIS) Ezr, 94


Cytosol, 3, 12, 106, 112 Ezrin, 1, 14, 15, 183

D F
Demyelination, 91, 93 Fasudil, 56
Dendritic cells, 36, 39, 74, 143, 144, 200, 261 Fatty acid, 126, 127, 133
Dermis, 70, 261 FDA, 51, 69, 78, 160, 192, 230,
Dexamethasone, 230 260–262, 264, 265
Diabetic nephropathy, 159, 163–165, 169 Fedratinib, 188, 192, 199, 261
Differentiation, 1, 3, 8, 9, 23, 40, 41, 52, 55, 56, 70, 73–75, FERM domain, 1, 11, 14–15, 183, 195
84, 93, 107, 109, 113, 125, 128, 148–150, 166, Fibrosis, 21, 42, 109–112, 114, 115, 133, 143–151, 159,
177, 189, 225, 227, 232, 243, 245, 247, 249 168, 169, 231
Diffuse large B-cell lymphoma (See DLBCL) Filaggrin, 70–73, 75
DLBCL, 193–199, 228, 229, 232 Filgotinib, 163, 260, 261
DNA binding domain, 2, 11, 16, 51, 110, 197 FIZZ.1, 55
DNM1L, 246 FLT3, 179, 181, 183, 225, 231, 232
DNMT1, 36, 38, 39 Flt-3 receptor, 242
Dopaminergic neuron, 87, 88 Fludarabine, 21
Down syndrome, 84, 93–94, 180, 228 FMS-like tyrosine kinase, 3 (See Flt-3)
DRP1, 246 Follicular lymphoma, 197, 228, 247
Dupilimab, 55 Food and Drug Administration (See FDA)
Four-point-one, 1
E Foxp2, 38
Foxp3, 21, 242, 243
E3 ligase complex (See ECS) Frameshift mutations, 7, 198, 245
EAE, 91–93 Fusion, 7, 179, 180–182, 186, 194–196, 225, 227, 228
EASI, 76, 77
Eccrine gland, 70
G
ECM, 143, 147
ECS, 20 G Protein-coupled receptor, 104, 115
Eczema herpeticum, 70 Gain-of-function, 7, 182, 183, 189, 190, 199, 226,
EGF, 105 229, 246
EGFR, 149, 165, 168, 169 Gamma-activated site (See GAS)
ELISA, 40 Gandotinib, 261
Elongin B, 6, 20 GAS, 5, 6, 242
Elongin BC, 6 GATA3, 55, 56, 57, 179, 243, 245
Elongin C, 6, 20 G-CSF, 4, 7, 232
Endogenous, 42, 75, 88, 104, 105, 110, 111, 166, 188 Gene, 2, 5–9, 16–19, 22, 23, 35, 40–43, 50, 52–55, 57–58,
Eosinophilic esophagitis, 247 70, 71, 73, 74, 87, 90, 91, 93, 95, 104–106, 108,
Eotaxin, 56, 57, 150 110, 111, 113, 114, 115, 124, 125, 127, 134, 144,
Eotaxin-3, 55 147–149, 160, 164, 165, 179, 180, 184, 191,
Epidermal growth factor (See EGF) 193, 194, 196–198, 225, 227, 228, 241–243,
Epidermis, 70, 75 245, 246, 249, 259, 261
Epigenetics, 1, 37–40, 242 Gene Set Enrichment Analysis, 91
EPO, 3, 4, 7, 162, 165, 167, 181, 182, 187, 189 Germline mutations, 182, 187, 190, 245
Epstein Barr Virus, 42, 246 GH, 3, 4, 7, 127, 133, 149, 161
ERK1, 16 Givinostat, 185
ERK2, 16 Glia, 83
ERMS, 15 Gluconeogenic, 125, 134
Erythrocytes, 107, 246 Glucose, 44, 123–135, 164, 165
Erythrocytosis, 182, 189, 190 Glucose tolerance, 125, 129, 131, 132
Erythropoietin (See EPO) Glycyrrhiza uralensis, 58
Essential thrombocythemia (See ET) GM-CSF, 2, 4, 7, 85, 91–93, 107, 160–162, 225
ET, 7, 186–192, 226, 227 GM-CSF receptor, 91, 162
Ethnic, 39 Goblet cell, 55, 58
ETV6, 179–181, 183 gp120, 258
Exome, 191, 246 gp130, 22, 84, 106, 110, 162, 167, 246
Extra cellular matrix (See ECM) gp130 receptor, 110
EZH2, 192 Gpc6, 85
272 Index

Granulocyte colony-stimulating factor (See G-CSF) I


Granulocyte-macrophage colony-stimulating factor (See
IBD, 248, 260, 261
GM-CSF)
Ichthyosis vulgaris, 70, 73
Granzyme, 166
IDH, 192
Growth hormone (See GH)
Ifih1, 94
Guanine, 114, 187
IFN, 6, 10, 19, 42, 50, 53, 93, 146, 161, 166, 167, 229, 242,
Guanine exchange factor, 114
243, 245, 246, 247, 262
GWAS, 50, 246, 247
IFN- α, 2, 6, 7, 10, 37, 84, 104, 107, 108, 143, 144, 146,
148
H IFNAR, 90, 93
IFNAR1, 93
H5N1, 42
IFNAR2, 93
HAND, 263, 264
IFN-β, 2, 6, 84, 85, 92, 143, 144, 146, 148
HBV, 42, 143, 148
IFN-γ, 2, 6, 7, 16, 20, 21, 23, 43, 44, 52–54, 84, 85, 87–89,
HCV, 41, 143, 146, 148
91–94, 104, 107, 108, 127, 143, 144, 145, 147,
HDM, 52
148, 160, 165, 168, 243, 245–247
Heart failure with reduced ejection fraction (See HFrEF)
IFN-λ, 52, 144, 146
Helminth, 243
IGF-1, 39, 149, 189
Hematopoiesis, 3, 9, 18, 177, 187, 188, 190, 191
IGH, 180, 182
Hematopoietic stem cell (See HSC)
IGK, 182
Hep3B, 144
IKZF1, 179, 182, 192
Heparin, 56
IL-10, 2–4, 18, 38, 84, 85, 92, 111, 147, 162, 196, 199, 229,
Hepatic fibrosis, 42, 143, 144, 146, 147
232, 242, 246
Hepatic steatosis, 133, 134
IL-11, 4, 84, 225
Hepatitis C, 41, 42, 143, 247
IL-12, 2, 4, 23, 50, 52–54, 84, 85, 91, 92, 147, 148, 150,
Hepatocellular carcinoma, 38, 41, 42, 144, 151, 245
160, 162, 167, 243, 245, 246, 247
Hepatocyte, 41, 125, 131, 133, 134, 143–151
IL-13, 4, 23, 49, 50, 52, 53, 55–58, 72–75, 85, 92, 150, 151,
Hepatocyte growth factor, 149
197, 225, 243, 245
HepG2, 37, 42
IL-13Rα1, 57
Heterodimer, 1, 3, 11, 16, 59, 144, 180, 198, 241, 242
IL-13Rα2, 57
HFrEF, 110, 111
IL-15, 4, 55, 84, 160, 162, 225, 228, 245, 262
HIES, 50, 52, 246
IL-15R, 246
HIF-2α, 187
IL-17A, 21, 57, 243, 245
High fat diet, 109, 125, 127, 128, 130, 132, 133, 134
IL-17F, 243, 245
Highly active antiretroviral therapy, 262
IL-19, 3, 4, 84
Histone acetyltransferase, 242
IL-1α, 87, 88
Histone deacetylase, 185, 193, 243
IL-1β, 43, 45, 85, 87–89
HIV, 260, 262–265
IL-2, 2, 4, 7, 41, 50, 54, 55, 84, 85, 107, 108, 148, 149, 160,
HIV-1, 258, 262, 263
168, 225, 228, 243, 247
HJC0123, 21
IL-20, 3, 4, 84
HJC0152, 22
IL-21, 4, 21, 84, 108, 160, 162, 228, 230, 232, 243
HL, 193, 197, 199
IL-21R, 161, 246
HLA-AB*27, 260
IL-22, 3, 4, 21, 72, 84, 134, 143, 147, 162, 196, 243, 245,
HLA-Cw0602, 260
246, 261
HO-3867, 22
IL-22BP, 147
Hodgkin lymphoma (See HL)
IL-23, 4, 85, 91, 92, 162, 245, 246, 247
Homodimer, 3, 9, 11, 16, 59, 111, 242
IL-24, 3, 4
Homology domain, 2, 49, 180
IL-26, 3, 4
House dust mite (See HDM)
IL-27, 4, 40, 144, 162, 232
HSC, 143–145, 147, 149, 150, 151, 177, 189, 190
IL-28A (See IFN-λ)
HuH-7 cell, 37
IL-28B (See IFN-λ)
Huntingtin, 90
IL-29 (See IFN-λ)
Huntington’s disease, 84, 90
IL-2R, 161, 246
HUVEC, 44
IL-2Rβ, 183
Hyperdiploidy, 179
IL-2Rγ, 7, 245
Hyper-IgE Syndrome (See HIES)
IL-3, 2, 4, 104, 107, 149, 160, 189, 190, 191, 225
Hyperplasia, 39, 40, 58, 70, 72, 74, 189
IL-31, 4, 50, 72, 74
Hypertension, 103, 107, 108, 110–114,
IL-33, 72, 73, 74
166, 168
IL-35, 4, 148
Hypodiploidy, 179
IL-3R, 2
Index 273

IL-4R, 56, 161, 246 Itch, 69, 70, 72, 74, 76, 77, 78
IL-4Rα, 55, 57, 69 ITP, 229, 230
IL-5, 4, 49–56, 72–74, 104, 107, 149, 150, 225, 243, 245
IL-5R, 2 J
IL-6, 2, 4, 5, 23, 37, 40, 41, 44, 50, 84, 85, 87–93, 104–109,
111, 112, 126, 127, 130, 133, 143, 147, 162, 167, J6/JFH-1, 37, 41
184, 199, 225, 229, 232, 243, 246, 247, 259, 260, Jak, 1–2
261, 263 Jak1, 2–4, 7, 10, 13, 15, 18, 37, 40, 41, 43–45, 49–51, 55,
IL-7, 4, 84, 107, 160, 225, 228, 245, 259, 262 59, 84, 85, 89, 91, 103–107, 112, 127, 128, 131,
IL-7R, 7, 246 136, 144, 147, 148, 160, 162–165, 167, 177,
IL-8, 40, 90, 91 180, 183–185, 189, 194–196, 198, 225, 228,
IL-9, 4, 49, 50, 54, 74, 84, 160, 225, 228, 243, 245 229, 231, 232, 242, 245, 246, 248, 258–263
IL-9Rα, 183 Jak2, 1–4, 7, 10–15, 18, 19, 21, 37, 38, 43–45, 49, 50,
ILC, 244, 245, 249 54–55, 84, 85, 87, 89, 91–93, 103–107, 109,
ILC1, 245 112–115, 125, 127–129, 131, 133–136, 144,
ILC2, 59, 74, 245 147, 160–165, 167, 169, 177, 180–183, 185,
ILC3, 245 187, 189, 190–192, 194, 197, 198, 225–231,
Immune thrombocytopenic purpura (See ITP) 233, 242, 246–248, 258, 259, 261
Immunodeficiency, 7, 50, 54, 103, 160, 241–249 Jak3, 1–4, 7, 10, 12, 13, 18, 49, 50, 51, 54, 84, 85, 103, 106,
Immunoglobulin, 229, 230, 246 107, 124, 125, 127, 129, 131, 133, 144,
Importin, 11, 12 160–163, 165, 168, 177, 181, 183, 184, 194,
Importin α, 11, 12 198, 225, 228, 230, 242, 245–248, 258–263
Inducible, 2, 3, 5, 6, 19, 42, 88–90, 93, 110, 243 Jakinib, 95
Inflammation, 21, 36, 40, 43–45, 49–59, 69, 73–75, Janus kinase (See Jak)
78, 85, 87, 91, 95, 108, 110, 111, 130, 132, JH (See Homology domain)
143–145, 147–151, 159, 163, 165, 200, 245, Jmjd3, 94
247, 257–265 JNK, 16
Inflammatory bowel disease (See IBD) JTE-052, 75–78
Influenza A virus, 42 Juvenile arthritis, 248
Innate Lymphoid Cell (See ILC) Juxtamembrane, 183, 190
iNOS, 45, 85, 87–89, 93, 94
Insulin, 18, 39, 124–135, 149
K
Insulin like growth factor (See IGF)
Interferon (See IFN) K539I, 13
Interferon regulatory factor (See IRF) Karyotype, 180
Interferon stimulated gene (See ISG) KchAP (See PIAS3)
Interferon-stimulated response element (See ISRE) KDM4C, 197
Intracellular, 1, 2, 9, 12, 17, 40, 49, 50, 72, 84, 85, 87, 112, Keratinocyte, 71–75
124, 127, 129, 130, 131, 133, 134, 144, 164, 165, Keratinocyte/epidermal barrier dysfunction, 73
177, 243, 258 Keratosis pilaris, 70
Invasion, 22, 38, 149 Kinase, 1–3, 6, 7, 9–14, 16, 17, 19, 21, 23, 35, 37, 40, 49,
Involucrin, 73 50, 56, 58, 69, 75, 84, 87, 103–106, 111–114,
Ionised calcium-binding adaptor molecule, 1, 90 124, 143, 145, 159–160, 162, 163, 165, 177,
IPEX, 74, 246 178–180, 182–183, 194, 196, 226–229, 231,
IRAK1, 199, 231 232, 241, 246–248
IRAK4, 199 Kinase inhibitory region (See KIR)
IRF1, 21 Kinesin, 15
IRF4, 242 KIR, 7, 19, 20
IRF7, 90, 94, 146 KIRT, 15
IRF9, 41, 42, 146 Knockdown, 41, 44, 58, 132, 148, 196, 197
Ischemia, 44, 85, 105, 109–111, 114, 115, 150, 151 Knockout mice, 19, 106, 110–111, 125, 127–128, 131–132,
ISG, 144 134, 147–149
ISGF3, 42 Kupffer cells, 143, 147, 149
ISIS481464 (See AZD9150)
Isoforms, 15, 16, 18, 42, 104–106, 115, 124, 125, 128,
L
132, 245
ISRE, 6 LacCer, 93
ISS-610, 22 Lactosylceramide (See LacCer)
ISS-840, 21 LAIR1, 182
Itacitinib, 136 LC3-1, 43
274 Index

Leflunomide, 51, 58 miR-124-STAT3, 90


Leptin, 4, 7, 18, 84, 109, 126, 127, 132, 134, 135, 232 miR-125b-5p, 39
Lestaurtinib, 199 miR-1264, 36, 38, 39
let-7c, 39 miR-130a, 37, 41
let-7f, 39 miR-146b, 36, 39
let-7i, 39 miR-150, 36, 39
Leukemia, 2, 7, 22, 40, 107, 110, 149, 177, 178, 182–187, miR-155, 36–38, 41–44
195, 196, 226–232 miR-181a, 36, 39
Leukemia inhibitory factor, 4, 7, 107, 149 miR-196a, 38
Leukemia inhibitory factor receptor, 149 miR-196b, 38
Leukopenia, 78 miR-19a, 36–39
Ligand, 1, 3, 7, 11, 16, 21, 43, 49, 89, 106, 109, 111, 112, miR-203, 36, 37, 40, 42
124, 160, 167, 180, 199, 258 miR-21, 36, 39, 40
Linker, 2, 11, 13, 15, 16 miR-214, 39
Lipid droplet, 126 miR-22, 39
Lipopolysaccharide (See LPS) miR-29, 39
LLL12, 21 miR-300-5p, 39
Lnk, 182, 191 miR-30a, 39
Loricrin, 72 miR-324-5p, 37, 42
Losartan, 112 miR-347, 39
Loss-of-function, 7, 115, 160, 177, 182, 185, 197, 198, 246 miR-373, 37, 41
Loss-of-heterozygosity, 187 miR-409 (miR-409-3p), 38
Low density lipoprotein, 128 miR-499, 36, 39
LPS, 38, 40, 44, 53, 90, 144, 150 miR-874, 39
Lupus nephritis, 159, 165, 166, 169 miR-9, 37, 44
LY5, 23 Missense, 50, 183, 187, 191, 194, 198, 245
Lymphadenopathy, 247 MMP, 147, 149
Lymphocyte, 52, 71, 75, 150, 166, 168 MMP2, 147
Lymphoma, 7, 193–200, 228–232, 246, 247 MMP9, 45
Lymphomagenesis, 194, 197, 199 Moesin, 1, 14, 15, 183
Lymphopenia, 246 Momelotinib, 232, 248, 261
Lyn, 58 Monocyte, 39, 91–93, 109, 162, 247
Lyn kinase, 58 Monocyte chemoattractant protein-1 (See MCP-1)
Monocytosis, 228
M Monosomy, 179
Monotherapy, 185, 186
Macrophage, 56, 84, 91–93, 109, 145, 160, 165, 228 Mouse embryonic fibroblast, 149
MAP kinase (See MAPK) MPN, 7, 163, 177, 180, 184, 186–194, 200, 226–228, 230,
MAPK, 16, 43, 112, 182, 185, 189, 197 231, 233
Mast cells, 51, 52, 54, 55, 74 MPTP, 89
Matrix metalloproteinase (See MMP) MS (See Multiple sclerosis)
Mcl-1, 110 mTORC1, 109, 229
MCL-1, 34, 40 MUC2, 58
MCP-1, 56 MUC4, 58
MDS, 228, 231 MUC5AC, 58
Megakaryocytes, 167, 229, 246 Mucin, 58
Megf10, 85 Mucocutaneous candidiasis, 246, 247
Mepolizumab, 54 Mucus, 51, 55–59
Mertk, 85 Multiple myeloma, 229, 231
Metastasis, 22, 149, 232 Multiple sclerosis, 84, 91, 246, 247
Metastatic bladder carcinoma, 245 Mutations, 1, 7, 10, 13, 16, 50, 52, 55, 56, 70, 71, 73, 90,
Methacholine, 53, 56 110–113, 115, 127, 135, 160, 163, 180–199,
Methylation, 19, 38, 242 225–230, 232, 233, 241, 245–249, 260, 261
Microglia, 83–95, 115 Mycobacteria, 245, 246
MicroRNA, 35, 38, 44 Mycophenolate mofetil, 168
Migration, 9, 21–23, 35, 38, 43, 44, 52, 54, 109 MyD88, 195, 198, 199
miR-1, 39 Myelodysplastic syndromes (See MDS)
miR-101, 39 Myelofibrosis, 160, 162, 187, 189, 190, 191, 200, 230–232,
miR-122, 37, 41 248, 257, 259, 260, 261
miR-1225-3p, 37, 42 Myeloproliferative neoplasms (See MPN)
Index 275

Myocardial infarction, 39, 107, 110, 111 Osteopontin, 111


Myogenesis, 125, 129 OVA, 53, 56, 57
Myokine, 129
Myosin, 15 P
Myotube, 129, 130
p15INK4B, 150
p21, 145
N
p38 MAPK, 16
NADPH, 88, 114 p62, 43
NADPH oxidase, 88, 114 Pacritinib, 188, 192, 195, 199, 231, 232
NAFLD (See Non-alcoholic fatty liver disease) Paeoniflorin, 165
NASH, 143, 148 Paracrine, 85, 129, 225, 257, 259–261
Nasopharyngitis, 77, 78 Parakeratosis, 70, 71
Natural killer cells (See NK cells) Parkinson’s disease, 84, 87
Natural killer-cell/T-cell lymphoma PAX5-JAK2, 228
(See NKTCL) PB1, 42
NC/Nga mice, 75 PB2, 42
Neointima, 39, 40 PBMC, 36, 39, 148, 194, 198
Neoplasm, 228 PCM1-JAK2, 228
NET, 166, 167 PD-L1, 43, 44, 196, 197, 200
Neuroinflammation, 84–94 PD-L2, 197
Neutropenia, 78, 248 Pediatric, 177, 178, 179, 183, 184, 186, 188, 191–194,
Neutrophil, 107, 166, 260 196–198, 228
Neutrophil extracellular trap (See NET) Peficitinib, 163, 260
New castle Disease Virus, 42 Perforin, 166
New onset diabetes mellitus after transplant (See Perivascular, 71, 73, 75
NODAT) PERSIST-1, 231
NF-κB, 39, 109, 111, 183, 198, 199, 229 PERSIST-2, 231
NHL, 193, 194, 199, 228 PF-04965842, 262
Nippostrongylus brasiliensis, 245 PF-06263276, 262
Nitric oxide (See NO) PH, 14, 179–186, 191, 231
Nivolumab, 199 PHD2, 187
NK cell, 146, 161, 245 Philadelphia chromosome, 179
NKTCL, 194–197 Phosphatase, 9, 17, 18, 104, 145, 198
NLS, 6, 11, 12 Phospholipase C, 112
NO, 44, 88, 89, 93 Phosphorylation, 3, 9, 11, 13, 14, 16–19, 21–23, 40, 44,
NODAT, 168, 169 55, 56, 58, 59, 74, 75, 85, 87, 104–106, 108, 109,
Non-alcoholic fatty liver disease, 132, 143 111–115, 124, 127, 128, 130, 132, 134, 135,
Non-alcoholic steatohepatitis (See NASH) 144, 147, 148, 160, 162, 165, 167, 180, 181, 183,
Non-canonical, 49, 104, 242 184, 189, 191, 196–198, 229, 232, 242, 247,
Non-Hodgkin Lymphoma (See NHL) 248, 258–260
Non-transmembrane, 17 Phosphotyrosine, 6, 9, 11, 14, 16–18, 21, 104
NSAID, 257 PI3K, 40, 45, 109, 182, 183, 185, 189, 229
Nuclear export sequence, 6 PIAS, 1, 5, 6, 7, 8, 17, 18, 19,106, 243
Nuclear localization sequence (See NLS) PIAS1, 7, 18, 19, 243
Nuclear pore complex, 11 PIAS3, 7, 19, 243
Nucleus, 1, 3, 5, 6, 8, 9, 11, 12, 16, 19, 49, 51, 70, 72, 87, PIASx, 7, 243
104–106, 111, 113, 124, 135, 144, 159, 164, PIASγ, 243
165, 169, 225, 241, 242, 259 Pim1, 94
Pimozide, 23, 59, 232
O PINIT, 19
Placebo, 76, 77, 78, 165, 192, 230
Oclacitinib, 51, 75, 76, 78, 79, 248 Pleckstrin homology (See PH)
Oligodendrocytes, 83 PM-242H, 24
Olumiant (See Baricitinib) PM-43I, 24
Oncogene, 186, 225–227, 242 PM-73G, 22
Oncostatin M (See OSM) PM-86I, 24
OSM, 4, 7, 84, 127, 225 PMF, 7, 186, 187, 189–192, 226, 227
Osteoarthritis, 36, 40 Polcythemia vera (See PV)
Osteoblasts, 36, 40 Poly (IC), 42
276 Index

Polymorphisms, 50, 53, 54, 115, 128, 144, 160, 167, 179, 113, 115, 124, 131, 134, 135, 144, 147, 149,
189, 246, 261 160–163, 165, 167, 168, 177, 180, 182, 183,
POMC, 135 185, 189, 190, 196, 198, 199, 225–228, 232,
Post-translational modifications, 6, 14, 242 233, 241–243, 245, 246–248, 258, 259
Pravastatin, 21 Renin, 112
Primary immunodeficiency, 7, 245 Reperfusion, 105, 109–111, 150, 151
Primary mediastinal large B-cell lymphoma (PMBCL), RESPONSE-1, 230
194–198 RESPONSE-2, 230
Primary myelofibrosis (See PMF) Restenosis, 39
PRKAR1A, 36, 40 Retinal pigment epithelial cells, 37, 43
PRL, 2–4, 7, 127, 149 Rheumatoid arthritis, 21, 36, 40, 51, 159, 246, 247, 259,
PRN371, 199 260, 262
Pro-inflammatory, 38, 84, 87–95, 109, 145, 147, 148, 150, Rho kinase, 56, 114
151, 163, 166, 227, 259–262, 264 RhoA, 23, 51, 56, 58, 114
Prolactin (See PRL) RING-finger like domain (See RLD)
Proliferation, 1, 3, 8, 9, 22, 35, 38, 41, 45, 52, 54, 55, 84, RLD, 7
109, 125, 132, 143, 144, 149, 167, 177, 178, 181, RNA polymerase, 42
182, 185–189, 193, 194, 196–200, 225, 226, RORα, 245
229, 232, 245, 261, 262 Rorγt, 21, 52, 242, 245
Promoter, 2, 5, 6, 10, 17, 22, 38, 42, 54, 57, 58, 87, 91, 109, rs167769, 56
128, 132, 242 rs16833215, 54
Pro-opiomelanocortin (See POMC) rs324011, 56
Protein tyrosine phosphatase (See PTP) rs324015, 56
Pruritus, 70, 74, 77, 78 rs4853546, 54
Pseudo tyrosine kinase (See PTK) rs7180246, 56
Pseudokinase, 1, 13, 104, 180, 182, 183, 227, 246 rs7574865, 148
Psoriasis, 74, 159, 162, 246–248, 257, 259, rs925847, 54
261, 262 Ruxolitinib, 51, 76, 109, 136, 160, 162, 181, 182, 184–186,
Psoriatic arthritis, 257, 260, 261 188, 192, 193, 195, 199, 200, 228–231, 233,
PSTPIP2, 145 260–264
PTK, 12, 13, 49
PTP, 6, 17, 18 S
PTP1B, 6, 17, 18, 198, 243
PTP-BL, 6, 17 S100β, 85
PTPN1, 195, 198 S3I-1757, 21
PTPN2, 18, 181, 183, 195, 198 S3I-201, 21, 105, 106, 114, 147
PTPRD, 243 S3I-201.1066, 21
PTPRT, 243 S3I-M2001, 21
PV, 13, 186–192, 226, 230, 261 S503P, 56
SAP domain, 7, 19
Q Scaffold attachment factor A/B, Acinus, PIAS (See SAP
domain)
Q576R, 56 sCD14, 264
sCD163, 264
R Schistosoma mansoni, 150
SCID, 7, 50, 107, 245, 246
R683G, 180–182, 228 SCORAD, 76, 77
Radixin, 1, 14, 15, 183 SEC31A-JAK2, 194, 229
RANBP6, 197 Sendai virus, 42
RanGAP, 12 Ser727, 16, 104, 108, 114
RanGDP, 12 Serine, 7, 11, 16, 17, 104, 145, 194, 242
RanGTP, 12 SH2, 1–3, 6, 7, 9, 11, 13–19, 21–24, 49, 51, 58, 104, 105,
RANTES, 56 114, 196, 197, 232, 241, 243, 248
Rapamycin, 109 SH2B3, 181–183, 188, 191, 192
Ras, 112, 183, 185 SH3, 14, 18
Rat insulin promoter, 132 SH-4-54, 22, 23
Rbx1, 6 SHP1, 243
Real time PCR, 40 SHP2, 20, 243
Receptors, 1–8, 9, 11, 14–21, 38, 42, 49–51, 53, 55, 57, 73, shRNA, 148
74, 85, 87, 89, 91–94, 104–107, 109, 110, 112, Sideroblasts, 228
Index 277

Signal transducer and activator of transcription Stattic, 21


(See STAT) STHdh111/111, 90
SIMPLIFY-1, 232 STING, 242
SIMPLIFY-2, 232 Stratum corneum, 70–73
Sinopulmonary, 246 Streptozotocin, 165
siRNA, 56–58, 114, 150 Stroke, 44, 103, 108, 111, 114–115
Skeletal muscle, 124, 129, 130 Stromal cells, 40, 55, 229
SKM (See Skeletal muscle) STX-0119, 21
SLE (See Systemic lupus erythematosus) Substantia nigra pars compacta, 87
SMAD, 40, 147 Sumoylation, 242
Smooth muscle cells, 36, 38, 39, 55, 57, 58, 107, 109, 112 Suppressor of cytokine signaling (See SOCS)
SNP, 50, 52, 54, 148 Survivin, 21, 109
SOCS, 1, 5–8, 9, 15, 17, 19, 20, 37, 44, 92, 104, 106, 165, Systemic lupus erythematosus, 165, 246, 247,
167, 197, 243 257, 262
SOCS1, 6, 7, 19–21, 36–38, 41, 42, 44, 105, 145, 165, 195,
197, 198, 243 T
SOCS2, 20, 38, 243
SOCS3, 6, 7, 19, 20, 36–39, 41, 105, 106, 109, 110, 114, Tacrolimus, 76, 77, 168
145, 147, 165, 243 TAD, 2, 104
SOCS4, 19, 243 Talin, 15
SOCS5, 36, 37, 44, 243 T-ALL, 7, 178, 179, 181, 183, 184, 186, 194, 228
SOCS6, 243 T-bet, 21, 53, 242, 243
SOCS7, 19, 243 TCPTP, 243
SOX2, 42, 43 TEL-JAK2, 183, 228
Sparcl1, 85 Tetrahydrocurcumin, 56
Sphingosine 1-phosphate, 109 Tetramerization, 242
Spinal cord, 37, 42, 43, 84 Tfh, 243
Splenomegaly, 163, 189, 191, 192, 231 TGF-β, 40, 93, 144, 147, 149, 159, 164
Spongiosis, 70, 71, 75 Th1, 23, 50, 52, 73, 91, 92, 144, 148, 150, 162, 243, 247,
Sprague-Dawley rats, 147 260, 261
Src homology 2 (See SH2) Th17, 41, 51, 91–93, 162, 243, 246, 247, 261
SSBP2-JAK2, 228 Th2, 23, 24, 50, 51, 69, 72–75, 78, 79, 150, 243,
STA-21, 21 247, 261
Staphylococcus aureus, 70 Th9, 74, 243
STAT, 1–3, 5–7, 9–12, 16, 18, 20–24, 35, 37, 38, 40, 42–45, THADA, 182
49, 51, 58, 59, 84, 85, 103, 104, 106, 108, 113, Thrombocytopenia, 163, 229–231
115, 124–129, 131, 133, 135, 136, 144, 160, Thrombocytosis, 163, 187–191, 228
164, 165, 182, 189, 198, 225, 226, 232, 233, Thrombopoietin (See TPO)
241–243, 248, 249, 257, 259 Thrombopoietin receptor (See TPOR)
STAT1, 2–6, 10–12, 16, 18–21, 23, 36–38, 40–44, 50–52, Thrombospondin, 85
58, 84, 85, 88–94, 103–109, 112–115, 124, 128, Thymic stromal lymphopoeitin (See TSLP)
134, 135, 144–146, 149, 151, 165, 177, 183, TIMP-1, 147, 148
184, 189, 196, 225, 232, 242, 243, 246, 247, 259 Titin, 111
STAT2, 2, 4, 10, 11, 16, 36, 39, 42, 50, 51–52, 84, 88–90, TLR, 87
103, 104, 135, 146, 151, 177, 225, 242, 243, TLR2, 44
246, 259 TLR4, 44
STAT3, 2, 4, 10, 12, 16, 18, 19, 21–23, 36, 38–44, 50–52, TMC-264, 23
58, 59, 75, 84, 85, 87–92, 94, 103–112, 114, 115, TNF receptor
128–136, 145–151, 162, 165, 167, 177, 181, TNFR1, 165
184, 189, 191, 194–196, 199, 225, 228–230, TNFR2, 165
232, 233, 242, 243, 245–247, 249, 259, 261 TNF-α, 38–40, 43–45, 85, 87–89, 93, 94, 126, 147,
STAT4, 4, 5, 10, 18, 23, 50–54, 84, 91, 103, 104, 125, 260, 261
145–148, 151, 177, 225, 242, 243, 245, 247, 259 TNF-β, 87, 88
STAT5A, 10, 36, 84, 103, 128, 225, 245, 247 Tofacitinib, 23, 51, 76, 77, 160, 162, 163, 166–169, 199,
STAT5B, 10, 84, 103, 128, 136, 191, 225, 245, 247 247, 248, 260–263
STAT6, 2, 4, 5, 10, 11, 18, 23, 24, 50–52, 54–59, 74, 75, 84, Toll-like receptor (See TLR)
85, 89, 103, 104, 115, 125, 128, 129, 132, 134, TPO, 3, 4, 161–163, 181, 190, 191, 226, 230
135, 145, 150, 151, 177, 195–198, 225, 242, TPOR, 7
243, 245, 247 Transactivation, 2, 11, 16, 51, 241
STAT6VT, 74, 75 Transactivation domain (See TAD)
278 Index

Transcription, 11, 21, 38, 49, 84, 87, 90, 103, 167, 225, U
242, 245
Uba7, 94
Transcription factor, 6, 19, 37, 42, 44, 51–58, 90, 113, 135,
Ubiquitin, 6, 15, 19, 20, 243
146, 148, 196
Ulcerative colitis, 257, 260, 261
Translocate, 1, 3, 5, 9, 11, 12, 16, 49, 87, 104–106, 112,
Upadcitinib, 260–262
124, 144, 225, 241, 242, 259
Ursolic acid, 57
Transmembrane, 2, 17, 38, 58, 106, 111, 183, 232, 241
Transphosphorylate, 11
Transverse aortic constriction, 39, 111 V
Treg, 74, 243
V617F, 13, 226–231, 246
T-regulatory (See Treg)
Vasculopathy, 166–168
TRIM24, 90
VEGF, 54, 110
Ts1Cje mouse, 93, 94
Venetoclax, 186
Ts65Dn mouse, 93
VHL, 187
TSLP, 4, 49, 50, 55, 59, 72–75, 180, 185
Vimentin, 85
Tumor, 20, 21, 22, 37, 38, 40, 44, 54, 147, 150,
Viral infection, 3, 41, 42, 52, 144, 146, 180, 246, 247,
182, 196–198, 200, 232, 241, 243, 245,
257–265
246, 260
VR588, 51
TYK2, 2, 7, 10, 13, 14, 15, 18, 49, 50, 84, 85, 87, 91,
103, 104, 124, 125, 127–129, 131, 144,
160–163, 165, 177, 196, 198, 225, 242, 246, W
258, 261, 262
Waldenstrom’s macroglobulinemia, 199
Type 1 diabetes, 109
Western blotting, 40
Type 2 diabetes, 123, 163
WHI-P97, 51
Type I IFN, 19, 166, 167, 229, 242–247
White adipose tissue, 126
Type II IFN, 242, 246
Wnt, 19, 40
TyrA1, 51
WSXWS motif, 3
Tyrosine, 1–3, 6, 8–12, 14, 16, 17, 19–21, 23, 49, 50, 55,
84, 89, 90, 103–106, 112, 124, 128, 144, 160,
167, 179, 183, 184, 196, 227, 228, 231, 232, X
241–243, 246, 248
Xenograft, 22, 75, 182, 185
Tyrosine kinase (See Tyk)
Xerosis, 70
Tyrphostin, 165

You might also like