Signal Transduction - Principles Pathways and Processes
Signal Transduction - Principles Pathways and Processes
Signal Transduction - Principles Pathways and Processes
ADVANCED TEXTBOOKS
EDITED BY
Signal transduction / edited by Lewis C. Cantley, Harvard Medical School, Tony Hunter, Salk Institute
for Biological Studies, Richard Sever, Cold Spring Harbor Laboratory, Jeremy Thorner, University of
California at Berkeley.
p. cm.
Summary: “This textbook provides a comprehensive view of signal transduction, covering both the
fundamental mechanisms involved and their roles in key biological processes. It first lays out the basic
principles of signal transduction, explaining how different receptors receive information and transmit it
via signaling proteins, ions, and second messengers. It then surveys the major signaling pathways that
operate in cells, before examining in detail how these function in processes such as cell growth and
division, cell movement, metabolism, development, reproduction, the nervous system, and immune
function”–Provided by publisher.
Includes bibliographical references and index.
ISBN 978-0-87969-901-7 (hardback)
ISBN 978-1-621822-30-1 (ebook)
1. Cellular signal transduction. 2. Developmental biology. 3. Pathology, Molecular. I. Cantley, Lewis,
editor of compilation. II. Hunter, Tony, 1943- editor of compilation. III. Sever, Richard, editor of
compilation. IV. Thorner, Jeremy W., editor of compilation.
QP517.C45S534 2013
571.7′4--dc23
2013043753
All World Wide Web addresses are accurate to the best of our knowledge at the time of printing.
Authorization to photocopy items for internal or personal use, or the internal or personal use of specific
clients, is granted by Cold Spring Harbor Laboratory Press, provided that the appropriate fee is paid
directly to the Copyright Clearance Center (CCC). Write or call CCC at 222 Rosewood Drive, Danvers,
MA 01923 (978-750-8400) for information about fees and regulations. Prior to photocopying items for
educational classroom use, contact CCC at the above address. Additional information on CCC can be
obtained at CCC Online at www.copyright.com.
For a complete catalog of all Cold Spring Harbor Laboratory Press publications, visit our website at
www.cshlpress.org.
This book is dedicated to the memory of Tony Pawson (1952–
2013). Tony was a giant in the field of signal transduction, who
established principles of protein–protein interactions that have
profoundly influenced our understanding of signal transduction.
His enduring legacy will be the discovery that the Src homology
2 (SH2) domain of one protein can selectively interact with a
tyrosine residue in a second protein, once it is phosphorylated in
response to an upstream signal. This type of inducible protein–
protein interaction can link intracellular signals generated in
response to various upstream stimuli to downstream signaling
events. This insight was the basis for his enormously influential
idea that eukaryotic signaling systems involve modular and
combinatorial interaction domains that propagate signals
throughout the cell.
Contents
Preface
Foreword
Edmond Fischer
3 Second Messengers
Alexandra C. Newton, Martin D. Bootman, and John D. Scott
mTOR Signaling
Mathieu Laplante and David M. Sabatini
Calcium Signaling
Martin D. Bootman
Wnt Signaling
Roel Nusse
Hedgehog Signaling
Philip W. Ingham
Notch Signaling
Raphael Kopan
Immunoreceptor Signaling
Lawrence E. Samelson
17 Vertebrate Reproduction
Sally Kornbluth and Rafael Fissore
22 Outlook
Jeremy Thorner, Tony Hunter, Lewis C. Cantley, and Richard Sever
Index
Preface
JEREMY THORNER
RICHARD SEVER
TONY HUNTER
LEWIS C. CANTLEY
Foreword
EDMOND FISCHER
University of Washington
SECTION I
SUMMARY
1 INTRODUCTION
Cells within multicellular organisms need to communicate with each other to
coordinate their growth, migration, survival, and differentiation. They do so
by direct cell–cell contact and secretion or release of molecules that bind to
and activate receptors on the surface of or inside target cells. Such factors can
stimulate the producer cell itself (autocrine stimulation), cells in the
immediate vicinity (paracrine stimulation), or cells in distant organs
(endocrine stimulation). The signaling induced within target cells is important
during embryonic development, as well as in the adult, where it controls cell
proliferation, differentiation, the response to infection, and numerous
organismal homeostatic mechanisms.
Many cell-surface receptors contain an extracellular ligand-binding
region, a single transmembrane segment, and an intracellular effector region,
which may or may not have an associated enzyme activity. Some receptors
contain multiple subunits that together form the ligand-binding site. Others,
including those encoded by the largest gene family in the human genome,
consist of a polypeptide that spans the cell membrane seven times. Finally,
there are receptors that are located intracellularly and are activated by ligands
that cross the cell membrane, such as steroid hormones. Below, we describe
the major families of ligands and receptors and the signal transduction
mechanisms they activate.
2 CELL-SURFACE RECEPTORS
2.1 Receptors with Associated Protein Kinase Activity
Several types of cell-surface receptors contain or are associated with kinase
activities that respond to the binding of a ligand. Perhaps best understood are
receptors with intrinsic protein tyrosine kinase domains. This receptor
tyrosine kinase (RTK) family has more than 50 human members (Lemmon
and Schlessinger 2010). RTKs have important roles in the regulation of
embryonic development, as well as in the regulation of tissue homeostasis in
the adult. Each has an extracellular, ligand-binding region, which consists of
different combinations of various domains, such as Ig-like, fibronectin type
III, and cysteine-rich domains. This is linked to a single transmembrane
segment and an intracellular region that includes a tyrosine kinase domain.
Based on their structural features, RTKs can be divided into 20 subfamilies
(Fig. 1), a well-studied example being the epidermal growth factor (EGF)
receptors (EGFRs).
Figure 1. Receptor tyrosine kinase (RTK) families. The 20 subfamilies of human RTKs and their
characteristic structural domains are shown. The individual members of each family are listed below.
(From Lemmon and Schlessinger 2010; adapted, with permission.)
2.2 Ligands
Each of these different receptor types responds to a subfamily of structurally
similar ligands. The ligands are normally small monomeric, dimeric, or
trimeric proteins, often derived by proteolytic processing from larger
precursors, some of which are transmembrane proteins. There is not a strict
specificity in ligand–receptor interactions within the families; normally each
ligand binds to more than one receptor, and each receptor binds more than
one ligand. Although it is rare that ligands for completely different types of
receptors bind to kinase-associated receptors, examples do exist.4
Figure 4. Schematic illustration of different modes of dimerization of RTKs. (A) Some dimeric ligands,
such as nerve growth factor (NGF), bind to receptors in a symmetric manner, but the receptors do not
contact each other. (B) Other dimeric ligands, such as stem cell factor (SCF), also bind to RTKs in a
symmetric manner, but the receptor dimer is in addition stabilized by direct receptor–receptor
interactions. (C) In the case of fibroblast growth factor (FGF), a ternary complex involving the ligand,
the receptor, and heparin/heparin sulfate stabilizes the receptor dimer. (D) In the case of members of
the epidermal growth factor (EGF) receptor family such as ErbB, ligand binding induces a
conformational change in the extracellular domain of the receptor that promotes direct receptor–
receptor interactions. (From Lemmon and Schlessinger 2010; adapted, with permission.)
In addition to homodimerization of receptors, heterodimerization also
often occurs. This is particularly common among cytokine receptors; a
ligand-specific receptor subunit often interacts with one or more common
subunits, such as gp130, βc, or γc, to form a heterodimer, heterotrimer, or
even more complicated multimer (Wang et al. 2009). Similarly, members of
different RTK subfamilies often form heterodimers. For instance, one
member of the EGFR subfamily, ErbB2, cannot bind to ligand itself, but acts
in heterodimers with other members of the family (Yarden and Sliwkowski
2001). In the PDGFR family, different dimeric isoforms of the ligand induce
formation of different dimeric complexes of PDGFα and PDGFβ receptors
(Heldin 1995). Because the downstream signaling pathways that are activated
to a large extent depend on the specific docking of SH2- or PTB-domain-
containing signaling molecules (see below and Ch. 2 [Lee and Yaffe 2014]),
differences in the autophosphorylation patterns of homodimeric versus
heterodimeric receptor complexes will give rise to different combinatorial
signals.
Serine/threonine kinase receptors present another variation on this theme.
These receptors are activated by ligand-induced assembly of two type I and
two type II receptors into heterotetrameric receptor complexes, in which
constitutively active type II receptors phosphorylate type I receptors in
serine- and glycine-rich sequences just upstream of the kinase domains. In
the TGFβ type I receptor, this causes a change in conformation that prevents
the inhibitory interaction of the receptor with the immunophilin FK506-
binding protein (FKBP12) and activates its kinase (Kang et al. 2009).
Note that certain receptors are present in dimeric complexes even in the
absence of ligand; examples include the erythropoietin cytokine receptor, and
the insulin and insulin-like growth factor-1 (IGF1) receptors, RTKs that
actually occur as disulfide-bonded dimers. Here, ligand binding induces
conformational changes that lead to activation of the receptor-associated
kinase. Importantly, there are indications that reorientation of the intracellular
regions of the receptors relative to each other in the dimer is important for
their activation (Jiang and Hunter 1999).
The heterotrimeric G proteins that relay signals from GPCRs are associated
with the underside of the plasma membrane and are composed of an α subunit
and a βγ dimer. Agonist-activated GPCRs act as GEFs that catalyze the
exchange of GDP bound to the α subunit for GTP, causing release of Gβγ
(Ch. 2 [Lee and Yaffe 2014]). A single ligand-bound GPCR can activate
several G proteins, providing the first layer of signal amplification. The GTP-
bound Gα subunits and Gβγ subunits can then promote the activation of a
variety of downstream effectors, stimulating a network of signaling events
that is highly dependent on the G-protein-coupling specificity of each
receptor (Fig. 6). The human G-protein α subunits are encoded by 16 distinct
genes and can be divided into four subfamilies: Gαs (Gαs and Gαolf), Gαi
(Gαt, Gαgust, Gαi1-3, Gαo, and Gαz), Gαq (Gαq, Gα11, Gα14, and Gα15/16), and
Gα12 (Gα12 and Gα13) (Fig. 6) (Cabrera-Vera et al. 2003). A single GPCR
can couple to either one or more than one family of Gα subunits. Five
different β subunits and 12 γ subunits that form functional βγ dimers have
been described.
Figure 6. Regulation of classical second messenger systems and Ras and Rho GTPases by GPCRs.
Agonist-activated GPCRs promote the dissociation of GDP bound to the α subunit pf heterotrimeric G
proteins and its replacement by GTP. Gα and Gβγ subunits can then activate numerous downstream
effectors. The 16 human G protein α subunits can be divided into the four subfamilies shown, and a
single GPCR can couple to one or more families of Gα subunits. Downstream effectors regulated by
their targets include a variety of second messenger systems, as well as members of the Ras and Rho
families of small GTP-binding proteins, which, in turn, control the activity of multiple MAPKs,
including ERK, JNK, p38, and ERK5. G-protein-dependent activation of these by GPCRs and β-
arrestin-mediated G-protein-independent activation of ERK and JNK can have multiple effects in the
cytosol. MAPKs also translocate to the nucleus, where they regulate gene expression. Activation of the
PI3K–Akt and mTOR pathways plays a central role in the regulation of cell metabolism, migration,
growth, and survival by GPCRs.
GPCRs stimulate pathways that control cell migration, survival, and growth
in part by activating MAPKs, a group of highly related proline-targeted
serine/threonine kinases that link multiple cell-surface receptors to
transcription factors. MAPKs include ERK1/2, JNK1-3, p38α-δ, and ERK5
MAPKs (Gutkind 1998; p. 81 [Morrison 2012]).
The small GTPase Ras, tyrosine kinases, PI3Ks, PKCs, and arrestins can
act downstream from GPCRs to promote the activation of ERK1/2 in a cell-
specific fashion (for review, see Gutkind 2000). The JNK cascade is activated
downstream from the small G proteins Rac, Rho, and Cdc42 (Coso et al.
1995). Indeed, Rac and Cdc42 can mediate signaling of Gβγ dimers and
Gα12, Gα13, Gαq, and Gαi to JNK (Gutkind 2000; Yamauchi et al. 2000).
Many GPCRs coupled to Gi activate Rac and JNK through the direct
interaction of Gβγ subunits with the P-REX1/2 family of Rac-GEFs (Welch
et al. 2002; Rosenfeldt et al. 2004). Similarly, Gαq activates Rho GTPases,
and hence JNK, through direct interaction with p63-RhoGEF and Trio (Lutz
et al. 2007). Gα12 and Gα13 bind to and act on three GEFs—p115-RhoGEF,
PDZ-RhoGEF, and LARG—to promote the activation of Rho downstream
from GPCRs (Hart et al. 1998; Fukuhara et al. 2001). Additional GEFs can
also contribute to this network. How GPCRs activate p38 and ERK5 is much
less clear, but, in general, these MAPKs are activated primarily by Gαq,
Gα12/13, and Gβγ dimers (Gutkind 2000).
Although human cancer-associated viruses express constitutively active
viral GPCRs, emerging data from deep sequencing studies have revealed that
a large fraction of human malignancies harbour mutations in GPCRs and G-
protein α subunits (O’Hayre et al. 2013). This has increased the interest in the
molecular mechanisms by which G proteins and GPCRs control normal and
cancer cell growth. Recent findings suggest that although GPCRs can
stimulate multiple diffusible-second-messenger-generating systems, their
ability to promote normal and aberrant cell proliferation often relies on the
persistent activation of Rho GTPases and MAPK cascades based on the direct
interaction of Gα subunits with RhoGEFs. The MAPKs regulate the activity
of nuclear transcription factors and coactivators, such as Jun, Fos, and YAP
(Yu et al. 2012; Vaqué et al. 2013).
Activation of the PI3K–Akt and mTOR pathways plays a central role in
cell metabolism, migration, growth, and survival (Zoncu et al. 2011). PI3K
generates 3′-phosphorylated inositol phosphates that participate in activation
of the kinase Akt and mTOR, which relay downstream signals (p. 87
[Hemmings and Restuccia 2012] and p. 91 [Laplante and Sabatini 2012]).
PI3Kγ shows restricted tissue distribution and is activated specifically by
GPCRs by the direct interaction of its catalytic (p110γ) and regulatory (p101)
subunits with Gβγ subunits (Lopez-Ilasaca et al. 1997). PI3Kγ is involved in
the chemokine-induced migration of leukocytes and plays significant roles in
innate immunity (Costa et al. 2011). In cells lacking PI3Kγ expression,
GPCRs can use PI3Kβ to stimulate phosphatidylinositol-(3,4,5)-tris-
phosphate (PIP3) synthesis (Ciraolo et al. 2008).
GPCRs are best known for their ability to control the activity of adenylyl
cyclases, phosphodiesterases, phospholipases, ion channels, and ion
transporters. The rapid regulation of these classical diffusible-second-
messenger-generating systems and their direct molecular targets is now
believed to represent a subset of the extensive repertoire of molecular
mechanisms deployed by GPCRs in physiological and pathological contexts.
Our recently gained understanding of GPCR signaling circuitries, including
GEFs, Ras and Rho GTPases, MAPKs, PI3Ks, and their numerous
downstream cytosolic and nuclear targets, provide a more global view of the
general systems by which these receptors exert their numerous physiological
and pathological roles. Indeed, the final biological outcome of GPCR
activation most likely results from the integration of the network of GPCR-
initiated biochemical responses in each cellular context. A new, systems-level
understanding may provide a molecular framework for the development of
novel approaches for therapeutic intervention in some of the most prevalent
human diseases.
The TNF family ligands are also transmembrane proteins but have
extracellular carboxyl termini. Intriguingly, these can be shed following
cleavage by proteases, and some inhibit the effects of the membrane-bound
ligand (Suda et al. 1997). The ligands are characterized by a conserved TNF
homology domain that mediates receptor binding. An exception is nerve
growth factor (NGF), which in addition to binding to an RTK also binds to
p75, a member of the TNFR family. Several ligands of this family bind to
more than one receptor (Fig. 7).
Figure 8. The integrin family of cell adhesion receptors. Integrins are composed of a heterodimer of
two transmembrane α and β subunits. The 18 α subunits and eight β subunits can form at least 24
heterodimeric complexes displaying distinct binding specificity and signaling capacity. (Adapted from
Hynes 2002.)
Figure 9. Integrin-based cell adhesion and signaling. Integrin engagement at cell-matrix adhesions or
interaction with a repertoire of cell-surface ligands results in the rapid assembly of a multifunctional
protein network (Geiger and Yamada 2011) containing many cytoskeletal, adaptor, and signaling
proteins. This contributes to cell adhesion and activates multiple signaling events. The adhesive
properties of integrins are, in turn, regulated by a variety of signaling pathways; this is known as inside-
out signaling.
5 NUCLEAR RECEPTORS
Nuclear receptors (NRs) comprise a large superfamily of intracellular
transcription factors that can effect gene expression changes in response to a
wide variety of lipophilic ligands (p. 129 [Sever and Glass 2013]). In this
respect, they differ from most other receptors in that they do not reside in the
plasma membrane, which many of their ligands can cross. Typical NR
ligands include steroids, vitamins, dietary lipids, cholesterol metabolites, and
xenobiotics. Ligands for 24 NRs have been identified; the remaining family
members are considered orphan receptors (Mangelsdorf et al. 1995). There
are 48 NRs in humans (49 in mice and 18 in Drosophila). NRs are believed
to be only one of two transcription factor families that are metazoan specific
(King-Jones et al. 2005; Degnan et al. 2009). Because many are endocrine
hormone receptors, this suggests a potentially critical role in the evolution of
animal physiology. Hormonal NRs are typically classified by the type of
ligands to which they bind, whereas orphans have various different
abbreviations, indicating, for example, similarity to known receptors (e.g.,
estrogen-related receptor, ERR). Binding of their cognate ligands, in the most
simplistic scenario, either changes the cellular localization of the NR or its
interaction with repressive and activating cofactors in the nucleus. What
distinguishes NRs from other genres of receptors is the ability to mediate
transcription without intermediate signaling cascades. Instead, they directly
bind to target genes. Nuclear–cytoplasmic cycling of some NRs allows them
to have nongenomic effects and also to be targets of cytoplasmic signaling
cascades (Wehling 1997).
Figure 10. General structure and binding of nuclear receptors. (A) Domain organization of a typical
nuclear receptor. (B) Three modes of signal transduction: as monomers, heterodimers, and homodimers.
(From Sonoda et al. 2008: modified, with permission, © Elsevier.)
5.3.1 Homodimers
Type I and type III receptors bind to DNA as homodimers when bound to
their cognate ligands. Type I receptors include all steroid receptors, and their
response elements typically consist of two hexameric inverted (palindromic)
repeat half-sites, for example, AGAACA (the glucocorticoid response
element, GRE) or AGAACA (the estrogen response element, ERE). These
sequences can mediate either activation or repression in response to hormone.
The DBDs of steroid receptors are highly similar. Thus, the glucocorticoid
receptor (GR), the mineralocorticoid receptor (MR), the androgen receptor
(AR), and the progesterone receptor (PR) all bind to overlapping response
elements. Type III receptors are similar to type I receptors except that they
recognize direct repeats instead of inverted repeats.
5.3.2 Heterodimers
RXR forms heterodimers with various nonsteroidal members of the NR
superfamily that do not bind to HREs efficiently by themselves. Depending
on the type of NR, HREs have specific conformations. Unlike the type I
receptors mentioned above, type II receptors form retinoid-acid-related-
receptor (RXR)-containing heterodimers that recognize direct repeats of
AGGTCA or similar sequences separated by specific spacing (Umesono et al.
1991). The RAR- and peroxisome proliferator-activated-receptor-γ (PPARγ)-
containing dimers RAR–RXR and PPARγ-RXR bind to a direct repeat (DR-
1), whereas the vitamin D3 receptor (VDR)-, TR-, and RAR-containing
dimers VDR–RXR, TR–RXR, and RAR–RXR bind to DR-3, DR-4, and DR-
5, respectively (Evans 2005).
5.3.3 Monomers
Type IV NRs bind to just one AGGTCA site as monomers in a ligand-
independent manner. These receptors use an extended DBD. There are
currently three known types of monomer-binding receptors, which are
defined by unique sequences 5′ to the consensus site. REV-ERBs tend to bind
to A(A/T)CTAGGTCA sequences, whereas ERRs bind to TNAAGGTCA
sequences.
7 GASES
The ability to sense and adapt to changes in levels of diatomic gases (O2, NO,
and CO) is crucial for an organism’s survival. Like steroids, gases are able to
cross the plasma membrane and bind to intracellular targets. Signal
transduction via gases is predominantly mediated by heme-containing
proteins. Binding of the gas to the sensor proteins via the heme moiety can
affect their activity or stability. In the case of oxygen, the source of the gas is
the environment. In contrast, nitric oxide and carbon monoxide can be
generated in neighboring cells and function as bona fide paracrine signals
(through covalent and noncovalent binding).
8 CONCLUDING REMARKS
Signaling mechanisms controlling cell growth, migration, survival, and
differentiation involve a plethora of different types of ligands, including
proteins, peptides, and certain lipids, that bind to receptors at the cell surface,
and steroids and gases that can pass across the plasma membrane and bind to
intracellular receptors. Recent work has provided immense insights into how
receptors are activated after ligand binding, how they initiate signaling inside
the cell, and how signaling is terminated. A striking feature is that regulated
dimerization or oligomerization is involved in the control of many plasma
membrane receptors and intracellular signaling molecules. Moreover, signal
transduction often occurs by the reversible formation of complexes and
specific translocations of signaling molecules, rather than by random
diffusion events. Another striking feature is that many different types of
posttranslational modifications of proteins are involved in the regulation of
the activity, stability, and subcellular localization of signaling molecules.
Furthermore, different signaling inputs converge on a few well-conserved
intracellular pathways, several of which link to transcriptional regulation of
gene programs.
Whereas the initial phase of signal transduction often involves
amplification mechanisms, such as inhibition of phosphatases in conjunction
with activation of tyrosine kinases, subsequent phases contain different
negative-feedback and termination mechanisms, acting at many different
levels. There is also extensive cross talk between different signaling
pathways. An important question that remains to be answered is exactly how
specificity of signaling is achieved. Moreover, most of our knowledge about
receptor signal transduction mechanisms comes from studies of cultured
cells; much more work is needed before we understand signal transduction at
the organismal level.
REFERENCES
*Reference is in this book.
Aggarwal BB. 2003. Signalling pathways of the TNF superfamily: A double-edged sword. Nat Rev
Immunol 3: 745–756.
Amit I, Citri A, Shay T, Lu Y, Katz M, Zhang F, Tarcic G, Siwak D, Lahad J, Jacob-Hirsch J, et al.
2007. A module of negative feedback regulators defines growth factor signaling. Nat Genet 39:
503–512.
Anastasiadis PZ, Reynolds AB. 2001. Regulation of Rho GTPases by p120-catenin. Curr Opin Cell
Biol 13: 604.
Ancot F, Foveau B, Lefebvre J, Leroy C, Tulasne D. 2009. Proteolytic cleavages give receptor tyrosine
kinases the gift of ubiquity. Oncogene 28: 2185–2195.
Andl CD, Mizushima T, Nakagawa H, Oyama K, Harada H, Chruma K, Herlyn M, Rustgi AK. 2003.
Epidermal growth factor receptor mediates increased cell proliferation, migration, and aggregation
in esophageal keratinocytes in vitro and in vivo. J Biol Chem 278: 1824–1830.
Andreeva AV, Kutuzov MA, Voyno-Yasenetskaya TA. 2007. Scaffolding proteins in G-protein
signaling. J Mol Signal 2: 13.
Arkhipov A, Shan Y, Das R, Endres NF, Eastwood MP, Wemmer DE, Kuriyan J, Shaw DE. 2013.
Architecture and membrane interactions of the EGF receptor. Cell 152: 557–569.
Ballon DR, Flanary PL, Gladue DP, Konopka JB, Dohlman HG, Thorner J. 2006. DEP-domain-
mediated regulation of GPCR signaling responses. Cell 126: 1079–1093.
Barak Y, Nelson MC, Ong ES, Jones YZ, Ruiz-Lozano P, Chien KR, Koder A, Evans RM. 1999.
PPARγ is required for placental, cardiac, and adipose tissue development. Mol Cell 4: 585–595.
Benovic JL, Pike LJ, Cerione RA, Staniszewski C, Yoshimasa T, Codina J, Caron MG, Lefkowitz RJ.
1985. Phosphorylation of the mammalian β-adrenergic receptor by cyclic AMP-dependent protein
kinase. Regulation of the rate of receptor phosphorylation and dephosphorylation by agonist
occupancy and effects on coupling of the receptor to the stimulatory guanine nucleotide regulatory
protein. J Biol Chem 260: 7094–7101.
Berman DM, Gilman AG. 1998. Mammalian RGS proteins: Barbarians at the gate. J Biol Chem 273:
1269–1272.
Blumer JB, Smrcka AV, Lanier SM. 2007. Mechanistic pathways and biological roles for receptor-
independent activators of G-protein signaling. Pharmacol Ther 113: 488–506.
Bockaert J, Pin JP. 1999. Molecular tinkering of G protein-coupled receptors: An evolutionary success.
EMBO J 18: 1723–1729.
Bodrikov V, Sytnyk V, Leshchyns’ka I, den Hertog J, Schachner M. 2008. NCAM induces CaMKIIα-
mediated RPTPα phosphorylation to enhance its catalytic activity and neurite outgrowth. J Cell Biol
182: 1185–1200.
Bouschet T, Martin S, Henley JM. 2005. Receptor-activity-modifying proteins are required for forward
trafficking of the calcium-sensing receptor to the plasma membrane. J Cell Sci 118: 4709–4720.
Braga VMM, Yap AS. 2005. The challenges of abundance: Epithelial junctions and small GTPase
signalling. Curr Opin Cell Biol 17: 466.
Brown GC, Cooper CE. 1994. Nanomolar concentrations of nitric oxide reversibly inhibit
synaptosomal respiration by competing with oxygen at cytochrome oxidase. FEBS Lett 356: 295–
298.
Brzostowski JA, Kimmel AR. 2001. Signaling at zero G: G-protein-independent functions for 7-TM
receptors. Trends Biochem Sci 26: 291–297.
Burgess AW, Cho HS, Eigenbrot C, Ferguson KM, Garrett TP, Leahy DJ, Lemmon MA, Sliwkowski
MX, Ward CW, Yokoyama S. 2003. An open-and-shut case? Recent insights into the activation of
EGF/ErbB receptors. Mol Cell 12: 541–552.
Cabrera-Vera TM, Vanhauwe J, Thomas TO, Medkova M, Preininger A, Mazzoni MR, Hamm HE.
2003. Insights into G protein structure, function, and regulation. Endocr Rev 24: 765–781.
* Cantrell D. 2014. Signaling in lymphocyte activation. Cold Spring Harb Perspect Biol doi:
10.1101/cshperspect.a018788.
Carmeliet P, Lampugnani MG, Moons L, Breviario F, Compernolle V, Bono F, Balconi G, Spagnuolo
R, Oostuyse B, Dewerchin M, et al. 1999. Targeted deficiency or cytosolic truncation of the VE-
cadherin gene in mice impairs VEGF-mediated endothelial survival and angiogenesis. Cell 98:
147–157.
Cavallaro U, Dejana E. 2011. Adhesion molecule signalling: Not always a sticky business. Nat Rev Mol
Cell Biol 12: 189–197.
Chan FK, Chun HJ, Zheng L, Siegel RM, Bui KL, Lenardo MJ. 2000. A domain in TNF receptors that
mediates ligand-independent receptor assembly and signaling. Science 288: 2351–2354.
Chuang TT, Iacovelli L, Sallese M, De Blasi A. 1996. G protein-coupled receptors: Heterologous
regulation of homologous desensitization and its implications. Trends Pharmacol Sci 17: 416–421.
Ciraolo E, Iezzi M, Marone R, Marengo S, Curcio C, Costa C, Azzolino O, Gonella C, Rubinetto C,
Wu H, et al. 2008. Phosphoinositide 3-kinase p110β activity: Key role in metabolism and mammary
gland cancer but not development. Sci Signal 1: ra3.
Coso OA, Chiariello M, Yu J-C, Teramoto H, Crespo P, Xu N, Miki T, Gutkind JS. 1995. The small
GTP-binding proteins Rac1 and Cdc42 regulate the activity of the JNK/SAPK signaling pathway.
Cell 81: 1137–1146.
Costa C, Martin-Conte EL, Hirsch E. 2011. Phosphoinositide 3-kinase p110γ in immunity. IUBMB Life
63: 707–713.
Croft M. 2009. The role of TNF superfamily members in T-cell function and diseases. Nat Rev
Immunol 9: 271–285.
Daniel JM, Reynolds AB. 1999. The catenin p120ctn interacts with Kaiso, a novel BTB/POZ domain
zinc finger transcription factor. Mol Cell Biol 19: 3614–3623.
Danishpajooh IO, Gudi T, Chen Y, Kharitonov VG, Sharma VS, Boss GR. 2001. Nitric oxide inhibits
methionine synthase activity in vivo and disrupts carbon flow through the folate pathway. J Biol
Chem 276: 27296–27303.
Degnan BM, Vervoort M, Larroux C, Richards GS. 2009. Early evolution of metazoan transcription
factors. Curr Opin Genet Dev 19: 591–599.
De Vos AM, Ultsch M, Kossiakoff AA. 1992. Human growth hormone and extracellular domain of its
receptor: Crystal structure of the complex. Science 255: 306–312.
* Devreotes P, Horwitz AR. 2013. Signaling networks that regulate cell migration. Cold Spring Harb
Perspect Biol doi: 10.1101/cshperspect.a005959.
Dioum EM, Rutter J, Tuckerman JR, Gonzalez G, Gilles-Gonzalez MA, McKnight SL. 2002. NPAS2:
A gas-responsive transcription factor. Science 298: 2385–2387.
Dorsam RT, Gutkind JS. 2007. G-Protein-coupled receptors and cancer. Nat Rev Cancer 7: 79–94.
Dumstrei K, Wang F, Shy D, Tepass U, Hartenstein V. 2002. Interaction between EGFR signaling and
DE-cadherin during nervous system morphogenesis. Development 129: 3983–3994.
Endres NF, Das R, Smith AW, Arkhipov A, Kovacs E, Huang Y, Pelton JG, Shan Y, Shaw DE,
Wemmer DE, et al. 2013. Conformational coupling across the plasma membrane in activation of
the EGF receptor. Cell 152: 543–556.
Evans RM. 2005. The nuclear receptor superfamily: A rosetta stone for physiology. Mol Endocrinol
19: 1429–1438.
Faustman D, Davis M. 2010. TNF receptor 2 pathway: Drug target for autoimmune diseases. Nat Rev
Drug Discov 9: 482–493.
Fedor-Chaiken M, Hein PW, Stewart JC, Brackenbury R, Kinch MS. 2003. E-cadherin binding
modulates EGF receptor activation. Cell Commun Adhes 10: 105–118.
Feldmann M, Maini RN. 2001. Anti-TNFα therapy of rheumatoid arthritis: What have we learned?
Annu Rev Immunol 19: 163–196.
Flower DR. 1999. Modelling G-protein-coupled receptors for drug design. Biochim Biophys Acta 1422:
207–234.
Forman BM, Chen J, Blumberg B, Kliewer SA, Henshaw R, Ong ES, Evans RM. 1994. Cross-talk
among RORα1 and the Rev-erb family of orphan nuclear receptors. Mol Endocrinol 8: 1253–1261.
Fox NL, Humphreys R, Luster TA, Klein J, Gallant G. 2010. Tumor necrosis factor-related apoptosis-
inducing ligand (TRAIL) receptor-1 and receptor-2 agonists for cancer therapy. Expert Opin Biol
Ther 10: 1–18.
Frisch SM, Screaton RA. 2001. Anoikis mechanisms. Curr Opin Cell Biol 13: 555–562.
Fuhrmann G, Chung AC, Jackson KJ, Hummelke G, Baniahmad A, Sutter J, Sylvester I, Scholer HR,
Cooney AJ. 2001. Mouse germline restriction of Oct4 expression by germ cell nuclear factor. Dev
Cell 1: 377–387.
Fukuhara S, Chikumi H, Gutkind JS. 2001. RGS-containing RhoGEFs: The missing link between
transforming G proteins and Rho? Oncogene 20: 1661–1668.
Geiger B, Yamada KM. 2011. Molecular architecture and function of matrix adhesions. Cold Spring
Harb Perspect Biol 3: a005033.
Geiger B, Bershadsky A, Pankov R, Yamada KM. 2001. Transmembrane crosstalk between the
extracellular matrix and the cytoskeleton. Nat Rev Mol Cell Biol 2: 793–805.
* Green DR, Llambi F. 2014. Cell death signaling. Cold Spring Harb Perspect Biol doi:
10.1101/cshperspect.a006080.
Grosheva I, Shtutman M, Elbaum M, Bershadsky AD. 2001. p120 catenin affects cell motility via
modulation of activity of Rho-family GTPases: A link between cell–cell contact formation and
regulation of cell locomotion. J Cell Sci 114: 695–707.
Gumbiner BM. 2005. Regulation of cadherin-mediated adhesion in morphogenesis. Nat Rev Mol Cell
Biol 6: 622–634.
Gutkind JS. 1998. The pathways connecting G protein-coupled receptors to the nucleus through
divergent mitogen-activated protein kinase cascades. J Biol Chem 273: 1839–1842.
Gutkind JS. 2000. Regulation of mitogen-activated protein kinase signaling networks by G protein-
coupled receptors. Sci STKE 2000: re1.
Hanyaloglu AC, von Zastrow M. 2008. Regulation of GPCRs by endocytic membrane trafficking and
its potential implications. Annu Rev Pharmacol Toxicol 48: 537–568.
* Hardie DG. 2012. Organismal carbohydrate and lipid homeostasis. Cold Spring Harb Perspect Biol
4: a006031.
* Harrison DA. 2012. The JAK/STAT pathway. Cold Spring Harb Perspect Biol 4: a011205.
Hart MJ, Jiang X, Kozasa T, Roscoe W, Singer WD, Gilman AG, Sternweis PC, Bollag G. 1998. Direct
stimulation of the guanine nucleotide exchange activity of p115 RhoGEF by Gα13. Science 280:
2112–2114.
Heldin C-H. 1995. Dimerization of cell surface receptors in signal transduction. Cell 80: 213–223.
* Hemmings BA, Restuccia DF. 2012. PI3K-PKB/Akt pathway. Cold Spring Harb Perspect Biol 4:
a011189.
Hsu SY, Kudo M, Chen T, Nakabayashi K, Bhalla A, van der Spek PJ, van Duin M, Hsueh AJ. 2000.
The three subfamilies of leucine-rich repeat-containing G protein-coupled receptors (LGR):
Identification of LGR6 and LGR7 and the signaling mechanism for LGR7. Mol Endocrinol 14:
1257–1271.
Hu X, Lazar MA. 1999. The CoRNR motif controls the recruitment of corepressors by nuclear
hormone receptors. Nature 402: 93–96.
Hubbard KB, Hepler JR. 2006. Cell signalling diversity of the Gqα family of heterotrimeric G proteins.
Cell Signal 18: 135–150.
Huttenlocher A, Horwitz AR. 2011. Integrins in cell migration. Cold Spring Harb Perspect Biol 3:
a005074.
Hynes RO. 2002. Integrins: Bidirectional, allosteric signaling machines. Cell 110: 673–687.
* Ingham PW. 2012. Hedgehog signaling. Cold Spring Harb Perspect Biol 4: a011221.
Issemann I, Green S. 1990. Activation of a member of the steroid hormone receptor superfamily by
peroxisome proliferators. Nature 347: 645–650.
Jiang G, Hunter T. 1999. Receptor signaling: When dimerization is not enough. Curr Biol 9: R568–
R571.
* Julius D, Nathans J. 2012. Signaling by sensory receptors. Cold Spring Harb Perspect Biol 4:
a005991.
Jura N, Endres NF, Engel K, Deindl S, Das R, Lamers MH, Wemmer DE, Zhang X, Kuriyan J. 2009.
Mechanism for activation of the EGF receptor catalytic domain by the juxtamembrane segment.
Cell 137: 1293–1307.
Kallen JA, Schlaeppi JM, Bitsch F, Geisse S, Geiser M, Delhon I, Fournier B. 2002. X-ray structure of
the hRORα LBD at 1.63 Å: Structural and functional data that cholesterol or a cholesterol
derivative is the natural ligand of RORα. Structure 10: 1697–1707.
Kang JS, Liu C, Derynck R. 2009. New regulatory mechanisms of TGFβ receptor function. Trends Cell
Biol 19: 385–394.
Kim N, Stiegler AL, Cameron TO, Hallock PT, Gomez AM, Huang JH, Hubbard SR, Dustin ML,
Burden SJ. 2008. Lrp4 is a receptor for Agrin and forms a complex with MuSK. Cell 135: 334–342.
Kim NG, Koh E, Chen X, Gumbiner BM. 2011. E-cadherin mediates contact inhibition of proliferation
through Hippo signaling-pathway components. Proc Natl Acad Sci 108: 11930–11935.
King-Jones K, Charles JP, Lam G, Thummel CS. 2005. The ecdysone-induced DHR4 orphan nuclear
receptor coordinates growth and maturation in Drosophila. Cell 121: 773–784.
Kniazeff J, Prezeau L, Rondard P, Pin JP, Goudet C. 2011. Dimers and beyond: The functional puzzles
of class C GPCRs. Pharmacol Ther 130: 9–25.
Kobielak A, Fuchs E. 2004. α-Catenin: At the junction of intercellular adhesion and actin dynamics.
Nat Rev Mol Cell Biol 5: 614–625.
Kobilka BK. 2011. Structural insights into adrenergic receptor function and pharmacology. Trends
Pharmacol Sci 32: 213–218.
* Kopan R. 2012. Notch signaling. Cold Spring Harb Perspect Biol 4: a011213.
Kopan R, Ilagan MXG. 2009. The canonical Notch signaling pathway: Unfolding the activation
mechanism. Cell 137: 216–233.
* Kornbluth S, Fissore R. 2014. Vertebrate reproduction. Cold Spring Harb Perspect Biol doi:
10.1101/cshperspect.a006064.
Korshunova I, Mosevitsky M. 2010. Role of the growth-associated protein GAP-43 in NCAM-
mediated neurite outgrowth. Adv Exp Med Biol 663: 169–182.
Krylova IN, Sablin EP, Moore J, Xu RX, Waitt GM, MacKay JA, Juzumiene D, Bynum JM, Madauss
K, Montana V, et al. 2005. Structural analyses reveal phosphatidyl inositols as ligands for the NR5
orphan receptors SF-1 and LRH-1. Cell 120: 343–355.
* Laplante M, Sabatini DM. 2012. mTOR signaling. Cold Spring Harb Perspect Biol 4: a011593.
* Lee MJ, Yaffe MB. 2014. Protein regulation in signal transduction. Cold Spring Harb Perspect Biol
doi: 10.1101/cshperspect.a005918.
Legate KR, Fassler R. 2009. Mechanisms that regulate adaptor binding to β-integrin cytoplasmic tails. J
Cell Sci 122: 187–198.
Legate KR, Wickstrom SA, Fassler R. 2009. Genetic and cell biological analysis of integrin outside-in
signaling. Genes Dev 23: 397–418.
Lehmann JM, McKee DD, Watson MA, Willson TM, Moore JT, Kliewer SA. 1998. The human orphan
nuclear receptor PXR is activated by compounds that regulate CYP3A4 gene expression and cause
drug interactions. J Clin Invest 102: 1016–1023.
Lemmon MA, Schlessinger J. 2010. Cell signaling by receptor tyrosine kinases. Cell 141: 1117–1134.
Lengyel E, Sawada K, Salgia R. 2007. Tyrosine kinase mutations in human cancer. Curr Mol Med 7:
77–84.
Leushacke M, Barker N. 2011. Lgr5 and Lgr6 as markers to study adult stem cell roles in self-renewal
and cancer. Oncogene 31: 3009–3022.
Lim X, Nusse R. 2013. Wnt signaling in skin development, homeostasis, and disease. Cold Spring
Harb Perspect Biol 5: a008029.
Lin S-Y, Makino K, Xia W, Matin A, Wen Y, Kwong KY, Bourguignon L, Hung M-C. 2001. Nuclear
localization of EGF receptor and its potential new role as a transcription factor. Nat Cell Biol 3:
802–808.
Liongue C, Ward AC. 2007. Evolution of Class I cytokine receptors. BMC Evol Biol 7: 120.
Locksley RM, Killeen N, Lenardo MJ. 2001. The TNF and TNF receptor superfamilies: Integrating
mammalian biology. Cell 104: 487–501.
Loers G, Schachner M. 2007. Recognition molecules and neural repair. J Neurochem 101: 865–882.
Lopez-Ilasaca M, Li W, Uren A, Yu J-C, Kazlauskas A, Gutkind JS, Heidaran MA. 1997. Requirement
of phosphatidylinositol-3 kinase for activation of JNK/SAPKs by PDGF. Biochem Biophys Res
Commun 232: 273–277.
Lu TT, Repa JJ, Mangelsdorf DJ. 2001. Orphan nuclear receptors as eLiXiRs and FiXeRs of sterol
metabolism. J Biol Chem 276: 37735–37738.
Luttrell LM, Gesty-Palmer D. 2010. Beyond desensitization: Physiological relevance of arrestin-
dependent signaling. Pharmacol Rev 62: 305–330.
Lutz S, Shankaranarayanan A, Coco C, Ridilla M, Nance MR, Vettel C, Baltus D, Evelyn CR, Neubig
RR, Wieland T, et al. 2007. Structure of Gαq-p63RhoGEF-RhoA complex reveals a pathway for
the activation of RhoA by GPCRs. Science 318: 1923–1927.
Magalhaes AC, Dunn H, Ferguson SS. 2012. Regulation of GPCR activity, trafficking and localization
by GPCR-interacting proteins. Br J Pharmacol 165: 1717–1736.
Mangelsdorf DJ, Borgmeyer U, Heyman RA, Zhou JY, Ong ES, Oro AE, Kakizuka A, Evans RM.
1992. Characterization of three RXR genes that mediate the action of 9-cis retinoic acid. Genes Dev
6: 329–344.
Mangelsdorf DJ, Thummel C, Beato M, Herrlich P, Schutz G, Umesono K, Blumberg B, Kastner P,
Mark M, Chambon P, et al. 1995. The nuclear receptor superfamily: The second decade. Cell 83:
835–839.
May LT, Leach K, Sexton PM, Christopoulos A. 2007. Allosteric modulation of G protein-coupled
receptors. Annu Rev Pharmacol Toxicol 47: 1–51.
* McCaffrey LM, Macara IG. 2012. Signaling pathways in cell polarity. Cold Spring Harb Perspect
Biol 4: a009654.
Mege RM, Gavard J, Lambert M. 2006. Regulation of cell–cell junctions by the cytoskeleton. Curr
Opin Cell Biol 18: 541–548.
Miaczynska M, Pelkmans L, Zerial M. 2004. Not just a sink: Endosomes in control of signal
transduction. Curr Opin Cell Biol 16: 400–406.
Miranti CK, Brugge JS. 2002. Sensing the environment: A historical perspective on integrin signal
transduction. Nat Cell Biol 4: E83–E90.
Miyamoto S, Teramoto H, Coso OA, Gutkind JS, Burbelo PD, Akiyama SK, Yamada KM. 1995.
Integrin function: Molecular hierarchies of cytoskeletal and signaling molecules. J Cell Biol 131:
791–805.
Miyamoto S, Teramoto H, Gutkind JS, Yamada KM. 1996. Integrins can collaborate with growth
factors for phosphorylation of receptor tyrosine kinases and MAP kinase activation: Roles of
integrin aggregation and occupancy of receptors. J Cell Biol 135: 1633–1642.
Moncada S, Palmer RM. 1991. Inhibition of the induction of nitric oxide synthase by glucocorticoids:
Yet another explanation for their anti-inflammatory effects? Trends Pharmacol Sci 12: 130–131.
Moncada S, Palmer RM, Higgs EA. 1991. Nitric oxide: Physiology, pathophysiology, and
pharmacology. Pharmacol Rev 43: 109–142.
Moquin D, Chan FK-M. 2010. The molecular regulation of programmed necrotic cell injury. Trends
Biochem Sci 35: 434–441.
* Morrison DK. 2012. MAP kinase pathways. Cold Spring Harb Perspect Biol 4: a011254.
Moser M, Legate KR, Zent R, Fassler R. 2009. The tail of integrins, talin, and kindlins. Science 324:
895–899.
Moustakas A, Heldin C-H. 2009. The regulation of TGFβ signal transduction. Development 136: 3699–
3714.
Mu Y, Sundar R, Thakur N, Ekman M, Gudey SK, Yakymovych M, Hermansson A, Dimitriou H,
Bengoechea-Alonso MT, Ericsson J, et al. 2011. TRAF6 ubiquitinates TGFβ type I receptor to
promote its cleavage and nuclear translocation in cancer. Nat Commun 2: 330.
Nabhan JF, Pan H, Lu Q. 2010. Arrestin domain-containing protein 3 recruits the NEDD4 E3 ligase to
mediate ubiquitination of the β2-adrenergic receptor. EMBO Rep 11: 605–611.
Nelson WJ, Nusse R. 2004. Convergence of Wnt, β-catenin, and cadherin pathways. Science 303:
1483–1487.
* Newton K, Dixit VM. 2012. Signaling in innate immunity and inflammation. Cold Spring Harb
Perspect Biol 4: a006049.
* Newton AC, Bootman MD, Scott JD. 2014. Second messengers. Cold Spring Harb Perspect Biol doi:
10.1101/cshperspect.a005926.
Ni C-Y, Murphy MP, Golde TE, Carpenter G. 2001. γ-Secretase cleavage and nuclear localization of
ErbB-4 receptor tyrosine kinase. Science 294: 2179–2181.
Nimmerjahn F, Ravetch JV. 2008. Fcγ receptors as regulators of immune responses. Nat Rev Immunol
8: 34–47.
Nose A, Nagafuchi A, Takeichi M. 1988. Expressed recombinant cadherins mediate cell sorting in
model systems. Cell 54: 993–1001.
O’Hayre M, Vázquez-Prado J, Kufareva I, Stawiski EW, Handel TM, Seshagiri S, Gutkind JS. 2013.
The emerging mutational landscape of G proteins and G-protein-coupled receptors in cancer. Nat
Rev Cancer 13: 412–424.
Ozdamar B, Bose R, Barrios-Rodiles M, Wang HR, Zhang Y, Wrana JL. 2005. Regulation of the
polarity protein Par6 by TGFβ receptors controls epithelial cell plasticity. Science 307: 1603–1609.
Pardee KI, Xu X, Reinking J, Schuetz A, Dong A, Liu S, Zhang R, Tiefenbach J, Lajoie G, Plotnikov
AN, et al. 2009. The structural basis of gas-responsive transcription by the human nuclear hormone
receptor REV-ERBβ. PLoS Biol 7: e43.
Pawson T. 2004. Specificity in signal transduction: From phosphotyrosine–SH2 domain interactions to
complex cellular systems. Cell 116: 191–203.
Pece S, Gutkind JS. 2000. Signaling from E-cadherins to the MAPK pathway by the recruitment and
activation of epidermal growth factor receptors upon cell–cell contact formation. J Biol Chem 275:
41227–41233.
Pece S, Chiariello M, Murga C, Gutkind JS. 1999. Activation of the protein kinase Akt/PKB by the
formation of E-cadherin-mediated cell–cell junctions. Evidence for the association of
phosphatidylinositol 3-kinase with the E-cadherin adhesion complex. J Biol Chem 274: 19347–
19351.
Perissi V, Jepsen K, Glass CK, Rosenfeld MG. 2010. Deconstructing repression: Evolving models of
co-repressor action. Nat Rev Genet 11: 109–123.
Pierce KL, Premont RT, Lefkowitz RJ. 2002. Seven-transmembrane receptors. Nat Rev Mol Cell Biol
3: 639–650.
Qian X, Karpova T, Sheppard AM, McNally J, Lowy DR. 2004. E-cadherin-mediated adhesion inhibits
ligand-dependent activation of diverse receptor tyrosine kinases. EMBO J 23: 1739–1748.
Raiborg C, Stenmark H. 2009. The ESCRT machinery in endosomal sorting of ubiquitylated membrane
proteins. Nature 458: 445–452.
Rajagopal S, Rajagopal K, Lefkowitz RJ. 2010. Teaching old receptors new tricks: Biasing seven-
transmembrane receptors. Nat Rev Drug Discov 9: 373–386.
Red Brewer M, Choi SH, Alvarado D, Moravcevic K, Pozzi A, Lemmon MA, Carpenter G. 2009. The
juxtamembrane region of the EGF receptor functions as an activation domain. Mol Cell 34: 641–
651.
Reinking J, Lam MM, Pardee K, Sampson HM, Liu S, Yang P, Williams S, White W, Lajoie G,
Edwards A, et al. 2005. The Drosophila nuclear receptor e75 contains heme and is gas responsive.
Cell 122: 195–207.
Renauld J-C. 2003. Class II cytokine receptors and their ligands: Key antiviral and inflammatory
modulators. Nat Rev Immunol 3: 667–676.
Rodgers KR. 1999. Heme-based sensors in biological systems. Curr Opin Chem Biol 3: 158–167.
Rosenfeld MG, Lunyak VV, Glass CK. 2006. Sensors and signals: A coactivator/corepressor/epigenetic
code for integrating signal-dependent programs of transcriptional response. Genes Dev 20: 1405–
1428.
Rosenfeldt H, Vazquez-Prado J, Gutkind JS. 2004. P-REX2, a novel PI-3-kinase sensitive Rac
exchange factor. FEBS Lett 572: 167–171.
* Samelson LE. 2011. Immunoreceptor signaling. Cold Spring Harb Perspect Biol 3: a011510.
* Sassone-Corsi P. 2012. The cyclic AMP pathway. Cold Spring Harb Perspect Biol 4: a011148.
Schlee S, Carmillo P, Whitty A. 2006. Quantitative analysis of the activation mechanism of the
multicomponent growth-factor receptor Ret. Nat Chem Biol 2: 636–644.
Schneider P, Thome M, Burns K, Bodmer JL, Hofmann K, Kataoka T, Holler N, Tschopp J. 1997.
TRAIL receptors 1 (DR4) and 2 (DR5) signal FADD-dependent apoptosis and activate NF-κB.
Immunity 7: 831–836.
* Sever R, Brugge JS. 2014. Signal transduction in cancer. Cold Spring Harb Perspect Med doi:
10.1101/cshperspect.a006098.
* Sever R, Glass CK. 2013. Signaling by nuclear receptors. Cold Spring Harb Perspect Biol 5:
a016709.
Shattil SJ, Kim C, Ginsberg MH. 2010. The final steps of integrin activation: The end game. Nat Rev
Mol Cell Biol 11: 288–300.
Shenoy SK, Lefkowitz RJ. 2011. β-Arrestin-mediated receptor trafficking and signal transduction.
Trends Pharmacol Sci 32: 521–533.
Sonoda J, Pei L, Evans RM. 2008. Nuclear receptors: Decoding metabolic disease. FEBS Lett 582: 2–9.
Stehlin-Gaon C, Willmann D, Zeyer D, Sanglier S, Van Dorsselaer A, Renaud JP, Moras D, Schule R.
2003. All-trans retinoic acid is a ligand for the orphan nuclear receptor RORβ. Nat Struct Biol 10:
820–825.
Suda T, Hashimoto H, Tanaka M, Ochi T, Nagata S. 1997. Membrane Fas ligand kills human
peripheral blood T lymphocytes, and soluble Fas ligand blocks the killing. J Exp Med 186: 2045–
2050.
Suenaert P, Bulteel V, Lemmens L, Noman M, Geypens B, Van Assche G, Geboes K, Ceuppens JL,
Rutgeerts P. 2002. Anti-tumor necrosis factor treatment restores the gut barrier in Crohn’s disease.
Am J Gastroenterol 97: 2000–2004.
Sueyoshi T, Kawamoto T, Zelko I, Honkakoski P, Negishi M. 1999. The repressed nuclear receptor
CAR responds to phenobarbital in activating the human CYP2B6 gene. J Biol Chem 274: 6043–
6046.
Suyama K, Shapiro I, Guttman M, Hazan RB. 2002. A signaling pathway leading to metastasis is
controlled by N-cadherin and the FGF receptor. Cancer Cell 2: 301–314.
Takai Y, Miyoshi J, Ikeda W, Ogita H. 2008. Nectins and nectin-like molecules: Roles in contact
inhibition of cell movement and proliferation. Nat Rev Mol Cell Biol 9: 603–615.
Taussig R, Gilman AG. 1995. Mammalian membrane-bound adenylyl cyclases. J Biol Chem 270: 1–4.
Terrillon S, Bouvier M. 2004. Roles of G-protein-coupled receptor dimerization. EMBO Rep 5: 30–34.
Thibeault S, Rautureau Y, Oubaha M, Faubert D, Wilkes BC, Delisle C, Gratton JP. 2010. S-
Nitrosylation of β-catenin by eNOS-derived NO promotes VEGF-induced endothelial cell
permeability. Mol Cell 39: 468–476.
Thiery JP, Sleeman JP. 2006. Complex networks orchestrate epithelial–mesenchymal transitions. Nat
Rev Mol Cell Biol 7: 131.
Tsukazaki T, Chiang TA, Davison AF, Attisano L, Wrana JL. 1998. SARA, a FYVE domain protein
that recruits Smad2 to the TGFβ receptor. Cell 95: 779–791.
Tsukita S, Furuse M, Itoh M. 2001. Multifunctional strands in tight junctions. Nat Rev Mol Cell Biol 2:
285–293.
Umesono K, Murakami KK, Thompson CC, Evans RM. 1991. Direct repeats as selective response
elements for the thyroid hormone, retinoic acid, and vitamin D3 receptors. Cell 65: 1255–1266.
van Amerongen R, Mikels A, Nusse R. 2008. Alternative wnt signaling is initiated by distinct
receptors. Sci Signal 1: re9.
Vaqué JP, Dorsam RT, Feng X, Iglesias-Bartolome R, Forsthoefel DJ, Chen Q, Debant A, Seeger MA,
Ksander BR, Teramoto H, et al. 2013. A genome-wide RNAi screen reveals a Trio-regulated Rho
GTPase circuitry transducing mitogenic signals initiated by G protein-coupled receptors. Mol Cell
49: 94–108.
Verma A, Hirsch DJ, Glatt CE, Ronnett GV, Snyder SH. 1993. Carbon monoxide: A putative neural
messenger. Science 259: 381–384.
Vleminckx K, Vakaet L Jr, Mareel M, Fiers W, van Roy F. 1991. Genetic manipulation of E-cadherin
expression by epithelial tumor cells reveals an invasion suppressor role. Cell 66: 107–119.
Wang GL, Jiang BH, Rue EA, Semenza GL. 1995. Hypoxia-inducible factor 1 is a basic-helix–loop–
helix-PAS heterodimer regulated by cellular O2 tension. Proc Natl Acad Sci 92: 5510–5514.
Wang X, Lupardus P, Laporte SL, Garcia KC. 2009. Structural biology of shared cytokine receptors.
Annu Rev Immunol 27: 29–60.
Warnmark A, Treuter E, Wright AP, Gustafsson JA. 2003. Activation functions 1 and 2 of nuclear
receptors: Molecular strategies for transcriptional activation. Mol Endocrinol 17: 1901–1909.
Wedel B, Humbert P, Harteneck C, Foerster J, Malkewitz J, Bohme E, Schultz G, Koesling D. 1994.
Mutation of His-105 in the β1 subunit yields a nitric oxide–insensitive form of soluble guanylyl
cyclase. Proc Natl Acad Sci 91: 2592–2596.
Wehling M. 1997. Specific, nongenomic actions of steroid hormones. Annu Rev Physiol 59: 365–393.
Wehrle-Haller B. 2012. Assembly and disassembly of cell matrix adhesions. Curr Opin Cell Biol 24:
569–581.
Wehrman T, He X, Raab B, Dukipatti A, Blau H, Garcia KC. 2007. Structural and mechanistic insights
into nerve growth factor interactions with the TrkA and p75 receptors. Neuron 53: 25–38.
Welch HC, Coadwell WJ, Ellson CD, Ferguson GJ, Andrews SR, Erdjument-Bromage H, Tempst P,
Hawkins PT, Stephens LR. 2002. P-Rex1, a PtdIns(3,4,5)P3- and Gβγ-regulated guanine-nucleotide
exchange factor for Rac. Cell 108: 809–821.
Whistler JL, Enquist J, Marley A, Fong J, Gladher F, Tsuruda P, Murray SR, Von Zastrow M. 2002.
Modulation of postendocytic sorting of G protein-coupled receptors. Science 297: 615–620.
Willard FS, Kimple RJ, Siderovski DP. 2004. Return of the GDI: The GoLoco motif in cell division.
Annu Rev Biochem 73: 925–951.
* Wrana JL. 2013. Signaling by the TGFβ superfamily. Cold Spring Harb Perspect Biol 5: a011197.
Yamauchi J, Kawano T, Nagao M, Kaziro Y, Itoh H. 2000. Gi-dependent activation of c-Jun N-
terminal kinase in human embryonal kidney 293 cells. J Biol Chem 275: 7633–7640.
Yanagisawa M, Anastasiadis PZ. 2006. p120 catenin is essential for mesenchymal cadherin-mediated
regulation of cell motility and invasiveness. J Cell Biol 174: 1087–1096.
Yarden Y, Sliwkowski MX. 2001. Untangling the ErbB signalling network. Nat Rev Mol Cell Biol 2:
127–137.
Yoshimura A, Mori H, Ohishi M, Aki D, Hanada T. 2003. Negative regulation of cytokine signaling
influences inflammation. Curr Opin Immunol 15: 704–708.
Yu FX, Zhao B, Panupinthu N, Jewell JL, Lian I, Wang LH, Zhao J, Yuan H, Tumaneng K, Li H, et al.
2012. Regulation of the Hippo–YAP pathway by G-protein-coupled receptor signaling. Cell 150:
780–791.
Yuzawa S, Opatowsky Y, Zhang Z, Mandiyan V, Lax I, Schlessinger J. 2007. Structural basis for
activation of the receptor tyrosine kinase KIT by stem cell factor. Cell 130: 323–334.
Zhang J, Huang W, Chua SS, Wei P, Moore D.D. 2002. Modulation of acetaminophen-induced
hepatotoxicity by the xenobiotic receptor CAR. Science 298: 422–424.
Zhang X, Gureasko J, Shen K, Cole PA, Kuriyan J. 2006. An allosteric mechanism for activation of the
kinase domain of epidermal growth factor receptor. Cell 125: 1137–1149.
Zikherman J, Weiss A. 2009. Antigen receptor signaling in the rheumatic diseases. Arthritis Res Ther
11: 202.
Zoncu R, Efeyan A, Sabatini DM. 2011. mTOR: From growth signal integration to cancer, diabetes and
ageing. Nat Rev Mol Cell Biol 12: 21–35.
Zwang Y, Yarden Y. 2009. Systems biology of growth factor-induced receptor endocytosis. Traffic 10:
349–363.
4The Ryk and Ror families of RTK, for example, bind members of the Wnt family. Ryk has a Wnt
inhibitory factor-1 (WIF1) domain, and the two Ror receptors have cysteine-rich domains related to a
domain in the Frizzled family of serpentine receptors to which ligands of the Wnt family bind (van
Amerongen et al. 2008).
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a005900
CHAPTER 2
SUMMARY
Outline
1 Introduction
2 Posttranslational modifications and the regulation of protein activity
3 Regulation of protein–protein interaction
4 Regulation of protein location
5 Regulation of protein production
6 Protein degradation
7 Concluding remarks: What does the future hold?
References
1 INTRODUCTION
Signal transduction processes are, in many respects, protein-driven events.
One common way to describe these circuits involves designating different
proteins, or modular domains within an individual protein, as “readers,”
“writers,” and “erasers.” In this model, catalytic domains that add specific
posttranslational modifications, such as kinases and acetyltransferases, are
called writers because they leave behind a physical mark on the proteins they
act on. Conversely, phosphatases, deacetylases, and other enzymes that
remove these modifications are examples of erasers. In addition to writers
and erasers, there are also domains that bind to specific posttranslationally
modified or unmodified sequences of amino acids. These types of domains
read the sequences produced by writers and erasers and are therefore called
readers. These readers can be involved in protein–protein interactions or
interactions between two parts of the same protein. Other readers can bind
directly to specific phospho- and neutral lipids, or to specific ions, such as
calcium, rather than to amino acid sequences. The targets of these readers,
writers, and erasers are often short amino acid sequences (typically three to
15 amino acids in length) called motifs. Using these modular parts, cells
dynamically encode information in response to environmental, chemical, or
developmental stimuli, to transduce the signal (see Table 1).
Figure 1. Allosteric modulation of protein activity. (A) Cartoon example of protein allostery. In this
example, an allosteric-regulation site exists in a region of the protein that is spatially distinct from the
active site. Modulation of the allosteric site (through posttranslational modification or the binding of a
ligand, cofactor, or protein) causes changes in the active site. Importantly, these changes can be either
activating or inhibiting. (B) Allosteric regulation of dihydrofolate reductase (DHFR). (Left panels) Two
surface views of the DHFR enzyme, highlighting the active site (bound to folate), cofactor binding site
(bound to NADPH), and residues involved in allosteric communication between the two sites (shown in
blue). (Middle panels) A view of a slice through the protein core. (Right panels) A cartoon
representation of the slice mappings shown in B. (Figure generously supplied by Rama Ranganathan
and Cell Press.)
Figure 2. Mechanism of kinase activation. (A) Conformational changes in protein kinases upon
phosphorylation enhance their catalytic ability. The structure of ERK2, an MAPK, is shown in its
inactive nonphosphorylated state and its active phosphorylated state. The carboxy-terminal lobes of the
kinase in both states (brown and red, respectively) have been superimposed. Note that following
phosphorylation of the activation loop, there is a marked rotation and reorientation of the amino-
terminal lobe (yellow and purple, respectively), bringing key catalytic residues, including those present
in a critical α helix, αC, into position, converting the kinase into an active state that can now
phosphorylate downstream substrates. (B) Close-up of phosphorylation-induced conformational
changes in the activation loop. Two key residues in the activation loop of MAPKs, a threonine and a
tyrosine residue, separated by a singe amino acid (i.e., a TXY motif) can interact with a network of
surrounding arginine residues only when they are in their phosphorylated states. These interactions not
only shift the positions of the threonine and tyrosine residues themselves (curved arrows), but drag the
entire activation loop into a new conformation that communicates with the rest of the protein to move
the entire amino-terminal lobe relative to the carboxy-terminal lobe, as shown in A.
Figure 4. The GTPase cycle. G proteins can be small (Ras-like) or large (heterotrimeric). Depicted here
is the nucleotide cycle for heterotrimeric G proteins, but the reactions are conceptually the same for
small GTPases. G-protein-coupled receptors (GPCRs) bind extracellular ligands, and transmit signals to
intracellular G proteins. The ligand-bound receptor functions as a guanine-nucleotide exchange factor
(GEF), causing the Gα subunit to exchange GDP for GTP. GTP-bound Gα no longer interacts with the
Gβγ dimer, and both entities are free to interact with downstream effector proteins. Gα controls the
duration of the signal because it is a GTPase, whose activity can be stimulated by GTPase-activating
proteins (GAPs) such as RGS proteins.
3 REGULATION OF PROTEIN–PROTEIN
INTERACTION
3.2 Motifs
For readers, writers, and erasers to function together to form signaling
networks, they must operate on common short amino acid sequence motifs in
their targets. Most such motifs contain ∼15 amino acid residues and include
particular amino acids that confer specificity to the writers and readers. For
example, CDKs and MAPKs (writers) preferentially phosphorylate S-P and
T-P motifs, whereas Polo-box domains (readers) preferentially recognize
motifs with the consensus S-pS/pT-P (where pS and pT are phosphorylated
serine and threonine). Similarly, many tyrosine kinases (writers)
phosphorylate tyrosine residues that are surrounded by acidic amino acid
residues, whereas SH2 domains (readers) prefer bind to phosphotyrosine-
containing sequences that contain specific patterns of hydrophobic or small
amino acid residues carboxy terminal to the phosphotyrosine. Some domains
can bind multiple sequence motifs or even bind ligands in multiple
orientations, as is the case for SH3 domains and SUMO-SIM domains.
This type of limited overlap between the sequences that the writer
domains generate and those that the reader domains recognize gives rise to
the specificity we observe in signaling networks. That is, only a small
fraction of the targets of any particular writer are then recognized by a
particular reader. As mentioned above, not all motifs require posttranslational
modification (e.g., proline-rich sequences are recognized by SH3 domains,
WW domains, and EVH1 domains, among others), which leads to
competition between a motif and multiple readers, depending on localization
and timing. In contrast, some types of domains appear to recognize specific
combinations of posttranslational modifications. Some bromo domains, for
example, recognize specific phosphoacetyl-methyl motif combinations on
histones (Filippakopoulos et al. 2012), whereas 14–3–3 proteins are able to
recognize both phosphorylated and phosphoacetylated histone sequences
(Macdonald et al. 2005). A similar effect can be achieved by adjacent
protein–protein interaction domains that bind to different posttranslational
modifications. These domains or domain combinations function as readers of
a more complex code, essentially creating “AND gates,” which can be used
to dictate greater specificity and control, or functioning as integrators of
signaling from multiple pathways.
Because most motifs are defined by the primary structure of a protein
rather than a complicated 3D arrangement of noncontiguous elements, it is a
relatively straightforward process to go motif hunting using protein
sequences and bioinformatics search tools (Obenauer et al. 2003; Obenauer
and Yaffe 2004). The small, “portable” nature of these motifs means that they
have frequently moved around within the sequences of evolutionarily related
proteins. Thus, a motif can sometimes be seen in one part of a protein
sequence in a human protein and in another part of the sequence in a yeast
ortholog.
4.1 Compartmentalization
A common theme in signaling is modulating compartmentalization of
signaling proteins within the cell. This can dictate access to substrates or
environmental conditions that promote activation. Subcellular compartments,
particularly organelles and cytoskeletal structures, can be very dynamic. In
these cases, signaling can promote the formation of these compartments. The
spatial segregation that organelles provide represents an important layer of
regulation, allowing similar proteins to execute different functions in
different environments. For example, mitotic kinases like Plk1, Aurora B,
and Never in mitosis kinase 2 (Nek2) achieve substrate specificity through
nonoverlapping subcellular localization, despite sharing partially overlapping
substrate selection motifs (Alexander et al. 2011).
Such compartmentalization can be achieved in different ways. Some
proteins have short sequences (also called signal peptides) that function as
localization signals, allowing transport to a particular organelle. The best
recognized of these are nuclear localization signals (NLSs), which typically
feature one or more short sequences of positively charged lysine or arginine
residues (Hung and Link 2011). Similar signals have been identified for
many other organelles, including mitochondria, lysosomes, and peroxisomes.
For proteins with localization sequences, regulated localization can be
achieved through posttranslational modifications or conformational changes
that mask or unmask a signal peptide. Type I nuclear receptors, for example,
are typically retained in the cytoplasm in an inactive complex with HSP90
(see p. 129 [Sever and Glass 2013]; Ch. 1 [Heldin et al. 2014]). Ligand
binding induces a conformational change, dissociation of HSP90,
homodimerization, and active transport into the nucleus, promoting DNA
binding and gene expression. Phosphorylation of the MAPK ERK on an SPS
motif within the kinase-insert region results in recognition by the nuclear
transport receptor importin β7 (Chuderland et al. 2008). Following nuclear
import, ERK then phosphorylates nuclear substrates, including transcription
factors Fos, Myc, and Elk1 (see p. 81 [Morrison 2012]). For proteins without
localization sequences, protein–protein interactions or association with lipid
membranes can localize them to specific subcellular regions.
Figure 6. Common forms of protein lipidation. The table highlights four lipid moieties that are
commonly used to modify proteins posttranslationally or (in the case of myristoylation)
cotranslationally. These modifications differ in terms of their consensus sequences, position of the
modification, hydrophobicity, and mechanism of regulation.
6 PROTEIN DEGRADATION
Every protein has a baseline level of expression resulting from the
equilibrium between its synthesis and breakdown (see Box 1). A control
mechanism widely used in signal transduction is regulation of the rate of
protein degradation. Many methods exist for regulated degradation of
proteins, including lysosomal and ER-mediated degradation (Ciechanover
2012), but the mechanism that seems to be most important in cell signaling is
ubiquitin-dependent proteasomal degradation (Hershko and Ciechanover
1998). Ubiquitin is a small 76-amino-acid protein that can be conjugated
through its carboxyl terminus to lysine residues on target proteins or to other
ubiquitin molecules to form ubiquitin chains that serve as a marker for
various cellular functions. Through diversity in chain length or the orientation
of chain attachment, ubiquitin can regulate protein trafficking and protein–
protein interactions, or function as a signaling scaffold (Muratani and Tansey
2003; Kirkin and Dikic 2007; Walczak et al. 2012).
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a005918
CHAPTER 3
Second Messengers
SUMMARY
Second messengers are small molecules and ions that relay signals
received by cell-surface receptors to effector proteins. They include a
wide variety of chemical species and have diverse properties that allow
them to signal within membranes (e.g., hydrophobic molecules such as
lipids and lipid derivatives), within the cytosol (e.g., polar molecules
such as nucleotides and ions), or between the two (e.g., gases and free
radicals). Second messengers are typically present at low concentrations
in resting cells and can be rapidly produced or released when cells are
stimulated. The levels of second messengers are exquisitely controlled
temporally and spatially, and, during signaling, enzymatic reactions or
opening of ion channels ensure that they are highly amplified. These
messengers then diffuse rapidly from the source and bind to target
proteins to alter their properties (activity, localization, stability, etc.) to
propagate signaling.
Outline
1 Introduction
2 Cyclic nucleotides
3 Lipid and lipid-derived second messengers
4 Ions as intracellular messengers
5 Concluding remarks
References
1 INTRODUCTION
Signals received by receptors at the cell surface or, in some cases, within the
cell are often relayed throughout the cell via generation of small, rapidly
diffusing molecules referred to as second messengers. These second
messengers broadcast the initial signal (the “first message”) that occurs when
a ligand binds to a specific cellular receptor (see Ch. 1 [Heldin et al. 2014]);
ligand binding alters the protein conformation of the receptor such that it
stimulates nearby effector proteins that catalyze the production or, in the case
of ions, release or influx of the second messenger. The second messenger
then diffuses rapidly to protein targets elsewhere within the cell, altering the
activities as a response to the new information received by the receptor.
Three classic second messenger pathways are illustrated in Figure 1: (1)
activation of adenylyl cyclase by G-protein-coupled receptors (GPCRs) to
generate the cyclic nucleotide second messenger 3′-5′-cyclic adenosine
monophosphate (cAMP); (2) stimulation of phosphoinositide 3-kinase (PI3K)
by growth factor receptors to generate the lipid second messenger
phosphatidylinositol 3,4,5-trisphosphate (PIP3); and (3) activation of
phospholipase C by GPCRs to generate the two second messengers
membrane-bound messenger diacylglycerol (DAG) and soluble messenger
inositol 1,4,5-trisphosphate (IP3), which binds to receptors on subcellular
organelles to release calcium into the cytosol. The activation of multiple
effector pathways by a single plasma membrane receptor and the production
of multiple second messengers by each effector can generate a high degree of
amplification in signal transduction, and stimulate diverse, pleiotropic,
responses depending on the cell type.
Figure 1. Second messengers disseminate information received from cell-surface receptors. Indicated
are three examples of a receptor activating an effector to produce a second messenger that modulates
the activity of a target. On the right, binding of agonists to a GPCR (the receptor) can activate adenylyl
cyclase (the effector) to produce cAMP (the second messenger) to activate protein kinase A (PKA; the
target). On the left, binding of growth factors to a receptor tyrosine kinase (RTK; the receptor) can
activate PI3K (the effector) to generate PIP3 (the second messenger), which activates Akt (the target).
In the center, binding of ligands to a GPCR (receptor) activates phospholipase C (PLC; the effector) to
clear the phospholipid PIP2, to generate two second messengers, DAG and IP3, which activate protein
kinase C (PKC; the target) and release calcium from intracellular stores, respectively.
Second messengers fall into four major classes: cyclic nucleotides, such
as cAMP and other soluble molecules that signal within the cytosol; lipid
messengers that signal within cell membranes; ions that signal within and
between cellular compartments; and gases and free radicals that can signal
throughout the cell and even to neighboring cells. Second messengers from
each of these classes bind to specific protein targets, altering their activity to
relay downstream signals. In many cases, these targets are enzymes whose
catalytic activity is modified by direct binding of the second messengers. The
activation of multiple target enzymes by a single second messenger molecule
further amplifies the signal.
Second messengers are not only produced in response to extracellular
stimuli, but also in response to stimuli from within the cell. Moreover, their
levels are exquisitely controlled by various homeostatic mechanisms to
ensure precision in cell signaling. Indeed, dysregulation of the second
messenger output in response to a particular agonist can result in cell/organ
dysfunction and disease. For example, chronic exposure to cAMP in the heart
results in an uncontrolled and asynchronous growth of cardiac muscle cells
called pathological hypertrophy. This early stage heart disease presents as a
thickening of the heart muscle (myocardium), a decrease in size of the
chamber of the heart, and changes in contractility. Because such prolonged
exposure to second messengers has deleterious effects, specific enzymes,
channels, and buffering proteins exist to rapidly remove second messengers,
either by metabolizing them or sequestering them away from target
molecules.
2 CYCLIC NUCLEOTIDES
2.1 cAMP and a Major Target, PKA
The “fight or flight” mechanism, more accurately referred to as the adrenal
response, prepares the body for situations of extreme stress. Release of the
hormone epinephrine, also known as adrenaline, from the adrenal gland into
the blood rapidly triggers vital cellular and physiological reactions that
prepare the body for intense physical activity. One of the most important
effects is the breakdown of glycogen into glucose to fuel muscles (Sutherland
1972).
Earl Sutherland first showed that epinephrine bound to cell-surface
receptors stimulates membrane-associated adenylyl cyclases to synthesize the
chemical messenger cAMP. Martin Rodbell and Alfred Gillman then
discovered that G-protein subunits are the intermediates that shuttle between
receptors and a family of eight adenylyl cyclase isoforms in the plasma
membrane. Each G protein is a trimer consisting of Gα, Gβ, and Gγ subunits.
The Gα subunits in the G proteins Gs and Gi are distinct and provide the
specificity for activation and inhibition of adenylyl cyclase, respectively. Gs
and Gi can thus couple binding of ligands to GPCRs with either activation or
inhibition of adenylyl cyclase, depending on the receptor type. This increases
or reduces production of cAMP, which diffuses from the membrane into the
cell. Edwin Krebs and Edmund Fischer later found that a principal task of
cAMP is to stimulate protein phosphorylation (Fischer and Krebs 1955),
ultimately showing cAMP-dependent protein kinase (also known as protein
kinase A, PKA) is responsible (Krebs 1993; Gilman 1995).
PKA is the major target for cAMP (see Fig. 2) (p. 99 [Sassone-Corsi
2012]). It is a heterotetramer consisting of two regulatory (R) subunits that
maintain two catalytic (C) subunits in an inhibited state. When cAMP levels
are low, the PKA holoenzyme is dormant; however, when cAMP levels are
elevated, two molecules bind in a highly cooperative manner to each R
subunit, causing a conformational change that releases the active C subunits
(Taylor et al. 2012). In humans, four genes encode R subunits (RIα, RIβ,
RIIα, and RIIβ) and two genes encode C subunits (Cα and Cβ), combinations
of which are expressed in most, if not all, tissues. The C subunits of PKA
phosphorylate serine or threonine residues on target substrates, typically
within the sequence RRxS/TΦ, in which Φ is an aliphatic hydrophobic
residue or an aromatic residue (Kemp et al. 1976). Around 300–500 distinct
intracellular proteins can be phosphorylated by PKA in a typical cell. These
include glycogen phosphorylase as part of the fight-or-flight mechanism.
Active phosphorylase catalyzes the production of glucose 1-phosphate, a
metabolic intermediate that is ultimately converted into other modified sugars
that are used in various aspects of cellular catabolism and ATP production
(see Ch. 14 [Hardie 2012]). On its release from the PKA holoenzyme, the C
subunit of PKA can diffuse into the nucleus, where it phosphorylates
transcription factors such as the cAMP-response-element-binding protein.
cAMP can thereby ultimately influence transcriptional activation and
reprogramming of the cell.
Figure 2. (A) cAMP is the archetypical second messenger. Its levels increase rapidly following
receptor-mediated activation of adenylyl cyclase (AC), which catalyzes the conversion of adenosine
monophosphate (AMP) to cAMP. This small second messenger activates PKA at specific cellular
locations as a result of anchoring of PKA to A-kinase-anchoring proteins (AKAPs). In addition, cAMP
activates EPACs (exchange proteins directly activated by cAMP). (B) AC activity is controlled by the
opposing actions of the Gs and Gi proteins. The cAMP produced by AC activates PKA by binding to
its R subunit. This releases the C subunit. The signal can be terminated by the action of
phosphodiesterase (PDE) enzymes.
A key reason for using ions as messengers is speed of response. Cells use
energy to maintain gradients of ions across their lipid membranes. By
activating channels or transporters, cells can use the potential energy
established by the electrochemical gradient of an ion to rapidly generate a
cellular signal (Clapham 2007). Unlike other intracellular messengers, ionic
signals can be generated with no enzymatic steps. The speed of the response
depends on the rate at which the intracellular concentration of the ion changes
and the proximity of the ions to their cellular targets. For example, action
potentials cause the fast release of neurotransmitters at nerve terminals
because the cytosolic concentration of calcium ions just beneath the plasma
membrane increases from ∼100 nM to >10 µM within milliseconds (Berridge
2006). In situations in which ions have to diffuse further before encountering
their target(s), the response will be slower. As we discuss below, the ability
to generate different spatiotemporal patterns, such as waves and oscillations,
is another important advantage of ionic intracellular messengers.
4.1 Calcium
Calcium is an extremely versatile intracellular messenger that controls a wide
range of cellular functions by regulating the activity of a vast number of
target proteins. The cellular effects of calcium are mediated either by direct
binding to a target protein, or stimulation of calcium sensors that detect
changes in calcium concentration and then activate different downstream
responses (Berridge 2004). The multitude of sensors that mediate effects of
calcium can be characterized by the nature of their calcium-binding site(s).
The most common calcium-binding motifs are EF-hands and C2 domains.
Synaptotagmin and troponin C are examples of proteins with C2 domains and
EF-hands, respectively. Calcium sensors also act by recruiting a range of
intermediary effectors, such as calcium-sensitive enzymes—for example,
calcium/calmodulin-dependent protein kinases (CaMKs), calcineurin (also
known as protein phosphatase 2B), myosin light chain kinase, and
phosphorylase kinase (see Ch. 13 [Kuo and Ehrlich 2014]).
4.6 Magnesium
Magnesium can also be considered an intracellular messenger because its
concentration can change dynamically in response to cellular stimulation (Li
et al. 2011). Given that magnesium binds to nucleotides, oligonucleotides,
and hundreds of enzymes, it is reasonable to conclude that it is an
intracellular messenger in its own right. Of its potential binding sites, ATP is
particularly important. Cellular processes typically use ATP complexed with
magnesium as an energy source. Moreover, magnesium has been shown to
cause prolonged inhibition of potassium channels in neurons following
muscarinic acetylcholine receptor activation (an effect that is not mimicked
by calcium), thereby regulating neuronal excitability (Chuang et al. 1997).
In addition, magnesium deserves consideration because it influences the
effects of calcium. Indeed, magnesium and calcium are typically thought to
have antagonistic actions. Magnesium frequently inhibits the transport and
cellular activities of calcium and can prevent pathological consequences of
increases in calcium levels (Romani 2013). Like calcium, there is an
electrochemical gradient of magnesium across the plasma membrane that can
serve as a reservoir for signal generation. The extracellular concentrations of
magnesium and calcium are similar (1.1–1.5 mM), and magnesium can also
act as a “calcimimetic” (e.g., by binding to the calcium-sensing receptor), a
GPCR that has pleiotropic actions. Within cells, the magnesium
concentration (0.3–1.5 mM) is several orders of magnitude higher than that of
calcium. It has been suggested that mitochondria might serve as a store of
magnesium and that magnesium can potentially regulate cellular respiration
(Wolf and Trapani 2012). Moreover, magnesium in the mitochondrial matrix
inhibits permeability transition pore (PTP) activation, an increase in the
leakiness of the inner mitochondrial membrane that allows solutes <1500 Da
to pass, and can precipitate mitochondrial swelling, apoptosis, and cell death.
Calcium promotes PTP, so magnesium acts as a counteracting antagonist.
Like calcium, magnesium has a plethora of transport pathways. Of these,
TRPM6 and TRPM7 (members of the TRP family) are relatively well
understood. TRPM6 is restricted to kidney tubules and the intestinal
epithelium, and plays an important role in magnesium (re)adsorption
(defective TRPM6 function leads to hypomagnesemia), whereas TRPM7 is
ubiquitously expressed in mammals. Both TRPM6 and TRPM7 are
“chanzymes”: ion channels that incorporate a kinase domain. Interestingly,
TRPM7 is regulated by PIP2, the source of the calcium-mobilizing messenger
IP3. Moreover, it binds to PLC, the enzyme that hydrolyses PIP2.
Consequently, hormonal stimulation of cells will lead to both a calcium
signal (via IP3 production) and a change in magnesium flux (via loss of PIP2-
mediated regulation of TRPM7) (Langeslag et al. 2007). MagT1 is also
believed to be a key cellular magnesium channel. Like TRPM6 and TRPM7,
it is located on the plasma membrane of mammalian cells. However, whereas
TRPM6 and TRPM7 allow both calcium and magnesium fluxes, it is believed
that MagT1 is a specific pathway for magnesium. Individuals bearing
mutations in the MAGT1 gene show high levels of Epstein–Barr virus
infection and a predisposition to lymphoma (Chaigne-Delalande et al. 2013).
5 CONCLUDING REMARKS
Second messengers disseminate information received by cellular receptors
rapidly, faithfully, and efficaciously. They are small, nonprotein organic
molecules or ions that bind to specific target proteins, altering their activities
in a variety of ways that allow them to respond appropriately to the
information received by receptors.
A key advantage of second messengers over proteins is that, unlike
proteins, second messenger levels are controlled with rapid kinetics. Thus,
whereas it may take tens of minutes for the levels of a protein to increase
significantly, most second messenger levels increase within microseconds
(e.g., ions) to seconds (e.g., DAG), They are often produced from precursors
that are abundant in cells or released from stores that contain high
concentrations of the second messenger; so, their generation is not rate
limiting. Thus, when the appropriate signal is received, second messengers
are rapidly generated, diffuse rapidly, and alter target protein function highly
efficiently.
Second messengers vary significantly in size and chemical character:
from ions to hydrophilic molecules such as cyclic nucleotides to hydrophobic
molecules such as diacylglycerol. Moreover, the continuing discovery of new
second messengers is expanding the repertoire of molecules known to convey
information within the cell. Indeed, only very recently, cyclic guanosine
monophosphate-adenosine monophosphate was shown to be a second
messenger that is synthesized by the enzyme cGAS in response to HIV
infection and binds to and activates a protein called STING, leading to
induction of interferon (Wu et al. 2013). The ability to respond rapidly to
information thus depends on an expanding library of small molecules.
REFERENCES
*Reference is in this book.
Baumgartner HK, Gerasimenko JV, Thorne C, Ferdek P, Pozzan T, Tepikin AV, Petersen OH, Sutton
R, Watson AJ, Gerasimenko OV. 2009. Calcium elevation in mitochondria is the main Ca2+
requirement for mitochondrial permeability transition pore (mPTP) opening. J Biol Chem 284:
20796–20803.
Beavo JA, Brunton LL. 2002. Cyclic nucleotide research—Still expanding after half a century. Nat Rev
Mol Cell Biol 3: 710–718.
Belevych AE, Radwanski PB, Carnes CA, Gyorke S. 2013. “Ryanopathy”: Causes and manifestations
of RyR2 dysfunction in heart failure. Cardiovasc Res 98: 240–247.
Berridge MJ. 2004. Calcium signal transduction and cellular control mechanisms. Biochim Biophys
Acta 1742: 3–7.
Berridge MJ. 2006. Calcium microdomains: Organization and function. Cell Calcium 40: 405–412.
Berridge MJ. 2014. Calcium regulation of neural rhythms, memory and Alzheimer’s disease. J Physiol
592: 281–293.
Berridge MJ, Lipp P, Bootman MD. 2000. The versatility and universality of calcium signalling. Nat
Rev Mol Cell Biol 1: 11–21.
Bootman MD, Berridge MJ, Roderick HL. 2002. Calcium signalling: More messengers, more channels,
more complexity. Curr Biol 12: R563–R565.
Bootman MD, Fearnley C, Smyrnias I, MacDonald F, Roderick HL. 2009. An update on nuclear
calcium signalling. J Cell Sci 122: 2337–2350.
Bos JL. 2003. Epac: A new cAMP target and new avenues in cAMP research. Nat Rev Mol Cell Biol 4:
733–738.
Cantley LC. 2002. The phosphoinositide 3-kinase pathway. Science 296: 1655–1657.
Catterall WA. 2011. Voltage-gated calcium channels. Cold Spring Harb Perspect Biol 3: a003947.
Chaigne-Delalande B, Li FY, O’Connor GM, Lukacs MJ, Jiang P, Zheng L, Shatzer A, Biancalana M,
Pittaluga S, Matthews HF, et al. 2013. Mg2+ regulates cytotoxic functions of NK and CD8 T cells
in chronic EBV infection through NKG2D. Science 341: 186–191.
Chuang H, Jan YN, Jan LY. 1997. Regulation of IRK3 inward rectifier K+ channel by m1
acetylcholine receptor and intracellular magnesium. Cell 89: 1121–1132.
Clapham DE. 2007. Calcium signaling. Cell 131: 1047–1058.
Colon-Gonzalez F, Kazanietz MG. 2006. C1 domains exposed: From diacylglycerol binding to
protein–protein interactions. Biochim Biophys Acta 1761: 827–837.
Fischer EH, Krebs EG. 1955. Conversion of phosphorylase b to phosphorylase a in muscle extracts. J
Biol Chem 216: 121–132.
Galione A. 2011. NAADP receptors. Cold Spring Harb Perspect Biol 3: a004036.
Gallegos LL, Newton AC. 2008. Spatiotemporal dynamics of lipid signaling: Protein kinase C as a
paradigm. IUBMB Life 60: 782–789.
Gees M, Owsianik G, Nilius B, Voets T. 2012. TRP channels. Compr Physiol 2: 563–608.
Gilman AG. 1995. Nobel Lecture. G proteins and regulation of adenylyl cyclase. Biosci Rep 15: 65–97.
Hannun YA, Obeid LM. 2008. Principles of bioactive lipid signalling: Lessons from sphingolipids. Nat
Rev Mol Cell Biol 9: 139–150.
* Hardie DG. 2012. Organismal carbohydrate and lipid homeostasis. Cold Spring Harb Perspect Biol
4: a006031.
* Heldin C-H, Lu B, Evans R, Gutkind JS. 2014. Signals and receptors. Cold Spring Harb Perspect
Biol doi: 10.1101/cshperspect.a005900.
* Hemmings BA, Restuccia DF. 2012. PI3K-PKB/Akt pathway. Cold Spring Harb Perspect Biol 4:
a011189.
Hogan PG, Lewis RS, Rao A. 2010. Molecular basis of calcium signaling in lymphocytes: STIM and
ORAI. Annu Rev Immunol 28: 491–533.
* Julius D, Nathans J. 2012. Signaling by sensory receptors. Cold Spring Harb Perspect Biol 4:
a005991.
Kemp BE, Benjamini E, Krebs EG. 1976. Synthetic hexapeptide substrates and inhibitors of 3′:5′-cyclic
AMP-dependent protein kinase. Proc Natl Acad Sci 73: 1038–1042.
* Kennedy MB. 2013. Synaptic signaling in learning and memory. Cold Spring Harb Perspect Biol doi:
10.1101/cshperspect.a016824.
Krebs EG. 1993. Nobel Lecture. Protein phosphorylation and cellular regulation I. Biosci Rep 13: 127–
142.
* Kuo IY, Ehrlich BE. 2014. Signaling in muscle contraction. Cold Spring Harb Perspect Biol doi:
10.1101/cshperspect.a006023.
Langeslag M, Clark K, Moolenaar WH, van Leeuwen FN, Jalink K. 2007. Activation of TRPM7
channels by phospholipase C-coupled receptor agonists. J Biol Chem 282: 232–239.
* Laplante M, Sabatini DM. 2012. mTOR signaling. Cold Spring Harb Perspect Biol 4: a011593.
Lemmon MA. 2007. Pleckstrin homology (PH) domains and phosphoinositides. Biochem Soc Symp
2007: 81–93.
Li FY, Chaigne-Delalande B, Kanellopoulou C, Davis JC, Matthews HF, Douek DC, Cohen JI, Uzel G,
Su HC, Lenardo MJ. 2011. Second messenger role for Mg2+ revealed by human T-cell
immunodeficiency. Nature 475: 471–476.
Manning BD, Cantley LC. 2007. AKT/PKB signaling: Navigating downstream. Cell 129: 1261–1274.
Newton AC. 2009. Lipid activation of protein kinases. J Lipid Res 50: S266–S271.
Nicoll DA, Ottolia M, Goldhaber JI, Philipson KD. 2013. 20 years from NCX purification and cloning:
Milestones. Adv Exp Med Biol 961: 17–23.
Nishizuka Y. 1992. Intracellular signaling by hydrolysis of phospholipids and activation of protein
kinase C. Science 258: 607–614.
Parekh AB. 2008. Ca2+ microdomains near plasma membrane Ca2+ channels: Impact on cell function.
J Physiol 586: 3043–3054.
Parys JB, De Smedt H. 2012. Inositol 1,4,5-trisphosphate and its receptors. Adv Exp Med Biol 740:
255–279.
Pitson SM. 2011. Regulation of sphingosine kinase and sphingolipid signaling. Trends Biochem Sci 36:
97–107.
Prins D, Michalak M. 2011. Organellar calcium buffers. Cold Spring Harb Perspect Biol 3: a004069.
Rizzuto R, Pozzan T. 2006. Microdomains of intracellular Ca2+: Molecular determinants and
functional consequences. Physiol Rev 86: 369–408.
Roderick HL, Berridge MJ, Bootman MD. 2003. Calcium-induced calcium release. Curr Biol 13:
R425.
Romani AM. 2013. Magnesium homeostasis in mammalian cells. Metal Ions life Sci 12: 69–118.
* Sassone-Corsi P. 2012. The cyclic AMP pathway. Cold Spring Harb Perspect Biol 4: a011148.
Schwaller B. 2010. Cytosolic Ca2+ buffers. Cold Spring Harb Perspect Biol 2: a004051.
Scott JD, Pawson T. 2009. Cell signaling in space and time: Where proteins come together and when
they’re apart. Science 326: 1220–1224.
Sutherland EW. 1972. Studies on the mechanism of hormone action. Science 171: 401–408.
Taylor SS, Ilouz R, Zhang P, Kornev AP. 2012. Assembly of allosteric macromolecular switches:
Lessons from PKA. Nat Rev Mol Cell Biol 13: 646–658.
Thul R, Coombes S, Roderick HL, Bootman MD. 2012. Subcellular calcium dynamics in a whole-cell
model of an atrial myocyte. Proc Natl Acad Sci 109: 2150–2155.
Tsui MM, York JD. 2010. Roles of inositol phosphates and inositol pyrophosphates in development,
cell signaling and nuclear processes. Adv Enzyme Regul 50: 324–337.
Vandecaetsbeek I, Vangheluwe P, Raeymaekers L, Wuytack F, Vanoevelen J. 2011. The Ca2+ pumps
of the endoplasmic reticulum and Golgi apparatus. Cold Spring Harb Perspec Biol 3: a004184.
Vaughan-Jones RD, Spitzer KW, Swietach P. 2009. Intracellular pH regulation in heart. J Mol Cell
Cardiol 46: 318–331.
Venteclef N, Jakobsson T, Steffensen KR, Treuter E. 2011. Metabolic nuclear receptor signaling and
the inflammatory acute phase response. Trends Endocrinol Metab 22: 333–343.
Wakai T, Vanderheyden V, Fissor RA. 2011. Ca2+ signaling during mammalian fertilization:
Requirements, players, and adaptations. Cold Spring Harb Perspect Biol 3: a006767.
Wolf FI, Trapani V. 2012. Magnesium and its transporters in cancer: A novel paradigm in tumour
development. Clin Sci (Lond) 123: 417–427.
Wong W, Scott JD. 2004. AKAP Signalling complexes: Focal points in space and time. Nat Rev Mol
Cell Biol 5: 959–971.
Wu J, Sun L, Chen X, Du F, Shi H, Chen C, Chen ZJ. 2013. Cyclic GMP-AMP is an endogenous
second messenger in innate immune signaling by cytosolic DNA. Science 339: 826–830.
Wymann MP, Schneiter R. 2008. Lipid signalling in disease. Nat Rev Mol Cell Biol 9: 162–176.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a005926
CHAPTER 4
SUMMARY
Outline
1 Introduction
2 Emergent properties of signaling networks
3 Information flow and processing
4 Concluding remarks
References
1 INTRODUCTION
Coordinated regulation of cellular processes allows cells to maintain
homeostatic balance and make decisions as to whether to divide, differentiate,
or die. In each case, the cell responds to chemical, mechanical, or electrical
signals, including hormones, neurotransmitters, mechanical stretch and shear,
and ion currents. The signaling pathways activate relay information to
effectors that alter subcellular processes, but they also process this
information. Information processing involves computation in which the
network of connected signaling molecules recognizes the amplitude and
duration of the incoming signal and produces an output signal of appropriate
strength and duration. This depends on the organization (topology) of the
signaling network and can change the relationship between inputs and
outputs. For example, the cell can limit responses to maintain homeostatic
balance or trigger a set of changes that takes it to another state (e.g.,
differentiation or division). The inherent computational abilities of signaling
networks thus provide the cell with decision-making capabilities.
In all signaling pathways, information flows through coupled biochemical
reactions and molecular interactions to the cellular machines that control its
output: biochemical (e.g., metabolic enzymes), mechanical (e.g., actin
cytoskeleton contractility), or electrical (e.g., ion channel activity in excitable
cells). Most of these pathways are part of wider networks (Weng et al. 1999).
Studying the flow of information (signal) through these networks (see Box 1)
is like studying traffic patterns, which depend on the population density of
the towns through which the highways pass, the state of local roads, the time
of the day, the weather, and other factors. Similarly, to understand the flow of
signals that regulate cellular machines, we must know how the pathways
involved are connected to form networks, the characteristics of the
connections, and how information is processed as it flows through these
interconnected pathways. Below, we discuss the organization of these
networks and explain how they can be modeled quantitatively.
Small G proteins, such as Ras, Rho, Rap, and Cdc42, are also major loci
of interconnectivity (Bar-Sagi and Hall 2000). Their guanine nucleotide
exchange factors (GEFs) and GTPase-activating proteins (GAPs) can be
regulated by numerous mechanisms, including binding of ligands such as
cAMP, calcium, and diacylglycerol (DAG), or posttranslational
modifications such as phosphorylation and acylation. Multiple receptors feed
into these GTPase regulators, and the small G proteins themselves can
modulate the activity of multiple signaling pathways by targeting several
different effectors (Fig. 1C).
Protein kinases and protein phosphatases are additional loci at which
multiple signaling pathways interact. Typically, a single protein kinase or
phosphatase can be activated by multiple receptors and has multiple
substrates. Together, all these components make the signaling pathways
within a cell extensively interconnected.
Most mammalian proteins are present as multiple isoforms. These can
arise from alternative splicing of a single mRNA or use of alternative
initiation codons, or they can be products of different genes altogether.
Typically, these isoforms have different characteristics that alter both
upstream and downstream interactions (connectivity) and/or their
intracellular localization. This further increases connectivity. An early
example is adenylyl cyclases (AC1-AC8), which allow signals from multiple
types of receptors and ion channels to feed into the cAMP pathway (Pieroni
et al. 1993; p. 99 [Sassone-Corsi 2012]). These enzymes can be activated or
inhibited by signals relayed via different types of receptors (Fig. 2), thus
enabling cAMP levels to provide an integrated measure of the information
coming into the cell from various sources. cAMP in turn works through
multiple effectors (e.g., protein kinase A, the EPAC Rap-GEF, and the cyclic-
nucleotide-gated channels) to regulate numerous cellular targets to evoke
physiological responses.
Figure 2. Different adenylyl cyclase (AC) isoforms are activated by multiple different upstream
signals. This allows the second messenger cAMP to provide an integrated measure of information from
several different sources.
1. Network Models
2. Dynamical Models
2.1 Ultrasensitivity
Sometimes a stimulus, such as receptor activation, can produce a switchlike
response in a downstream signaling component. In such a system, at a certain
point called the threshold, a small change in the ligand/receptor can cause a
large change in the activity of a downstream effector. Such responses are
called ultrasensitive. This was first observed with coupled enzymes switching
between activity states (Goldbeter and Koshland 1981). Subsequently, similar
ultrasensitivity has been observed in control of the cell cycle by the ERK
pathway, where small changes in stimulus cause large changes in ERK
activity (Ferrell and Machleder 1998). Ultrasensitivity can be produced by
several mechanisms such as cooperativity, multistep regulation as seen in the
ERK pathway, and by changing the levels of activators to inhibitors of a
signaling component such as a GTPase (Lipshtat et al. 2010).
2.2 Bistability
The ability of positive-feedback loops like the one involving PKC, ERK, and
PLA2 (Fig. 3C) to function as switches is an emergent property called
bistability. ODE-based models (see Box 2) of the ERK-PLA2-PKC pathway
(Bhalla and Iyengar 1999) allow one to plot the activity of ERK as a function
of protein kinase C activity and vice versa (Fig. 4). The two curves intersect
three times. In this system, if the initial stimulus is above the threshold level
(the middle intersection point) needed to increase activity of either protein
kinase C or ERK, the system will move from the lower intersection point
(basal state) to the upper intersection point (active state) for a prolonged
period. The upper and lower intersection points represent the two stable
states. The middle intersection point represents the level above which the
initial stimulus engages the feedback loop and moves the system to the
activated equilibrium state. Otherwise, when the stimulus is withdrawn the
system will revert to its basal equilibrium state. The computational modeling
analysis shows that the connectivity alone is insufficient for the switching
behavior. The concentrations of the signaling components and the reaction
rates are critical in ensuring that a feedback loop with this topology functions
as a bistable switch.
Figure 4. Activity states of components in a positive-feedback loop can be plotted as functions of each
other to study the characteristics of a bistable system that switches from an inactive basal state (state 1)
to an active state (state 2) once a threshold is reached. For example, when the concentration of active
MAPK or PKC is above the threshold levels upon growth factor receptor activation, the system moves
to state 2. The plot is obtained by computational modeling of the positive-feedback loop in Figure 3C,
and recording the steady-L-state levels of active PKC for a given fixed concentration of active MAPK
(black dashed line) upon stimulation, and vice versa (red solid line).
4 CONCLUDING REMARKS
Signaling networks in cells produce outputs that are manifested as decisions
to perform physiological functions in response to biochemical, electrical, or
mechanical stimuli. These outputs are most often controlled by protein
kinases in the cytoplasm and near the plasma membrane. These protein
kinases phosphorylate and regulate metabolic enzymes, channels,
transporters, and components of the cytoskeletal machinery. In the nucleus,
the targets are typically transcription factors and proteins that control
chromosomal organization and dynamics. The outputs, whether they are
production of glucose from glycogen by liver cells in response to
epinephrine, firing of neurons in response to neurotransmitters, or hormone-
driven changes in gene expression of ovarian cells, reflect both the
characteristics (amplitude and duration) of the external signals and
information processing within signaling networks. The ability to balance
these allows the cell to respond to varying stimuli and return to homeostatic
balance.
When signaling components are inappropriately activated or inactivated
(e.g., by mutations), however, the information-processing capability of the
networks is also altered. This may lead to sustained changes in gene
expression that push the cell into a different state, for example, an enhanced
rate of proliferation, which is often detrimental and can cause diseases such
as cancer (Ch. 21 [Sever and Brugge 2014]). Similarly, bacterial toxins can
cause disease by altering components within signaling networks (Ch. 20
[Alto and Orth 2012]). Both examples show how critical it is for us to
understand the organization and information-processing capability of
signaling networks.
REFERENCES
*Reference is in this book.
Alon U. 2007. Network motifs: Theory and experimental approaches. Nat Rev Genet 8: 450–461.
* Alto NM, Orth K. 2012. Subversion of cell signaling by pathogens. Cold Spring Harb Perspect Biol
4: a006114.
Bagci EZ, Vodovotz Y, Billiar TR, Ermentrout GB, Bahar I. 2006. Bistability in apoptosis: Roles of
bax, bcl-2, and mitochondrial permeability transition pores. Biophys J 90: 1546–1559.
Barabasi AL, Albert R. 1999. Emergence of scaling in random networks. Science 286: 509–512.
Bar-Sagi D, Hall A. 2000. Ras and Rho GTPases: A family reunion. Cell 103: 227–238.
Bhalla US, Iyengar R. 1999. Emergent properties of networks of biological signaling pathways. Science
283: 381–387.
Biggs N, Lloyd EK, Wilson RJ. 1986. Graph theory 1736–1936. Clarendon, Oxford.
* Bootman M. 2012. Calcium signaling. Cold Spring Harb Perspect Biol 4: a011171.
Brandman O, Ferrell JE Jr, Li R, Meyer T. 2005. Interlinked fast and slow positive feedback loops
drive reliable cell decisions. Science 310: 496–498.
* Duronio RJ, Xiong X. 2013. Signaling pathways that control cell proliferation. Cold Spring Harb
Perspect Biol 5: a008904.
Ferrell JE Jr, Machleder EM. 1998. The biochemical basis of an all-or-none cell fate switch in Xenopus
oocytes. Science 280: 895–898.
Gillespie DT. 2007. Stochastic simulation of chemical kinetics. Annu Rev Phys Chem 58: 35–55.
Goldbeter A, Koshland DE Jr. 1981. An amplified sensitivity arising from covalent modification in
biological systems. Proc Natl Acad Sci 78: 6840–6844.
* Green DR, Llambi F. 2014. Cell death signaling. Cold Spring Harb Perspect Biol doi:
10.1101/cshperspect.a006080.
* Heldin C-H, Lu B, Evans R, Gutkind JS. 2014. Signals and receptors. Cold Spring Harb Perspect
Biol doi: 10.1101/cshperspect.a005900.
* Hemmings BA, Restuccia DF. 2012. The PI3K-PKB/AKT pathway. Cold Spring Harb Perspect Biol
4: 011189.
Jordan JD, Landau EM, Iyengar R. 2000. Signaling networks: The origins of cellular multitasking. Cell
103: 193–200.
Lipshtat A, Purushothaman SP, Iyengar R, Ma’ayan A. 2008. Functions of bifans in context of multiple
regulatory motifs in signaling networks. Biophys J 94: 2566–2579.
Lipshtat A, Jayaraman G, He JC, Iyengar R. 2010. Design of versatile biochemical switches that
respond to amplitude, duration, and spatial cues. Proc Natl Acad Sci 107: 1247–1252.
Mangan S, Alon U. 2003. Structure and function of the feed-forward loop network motif. Proc Natl
Acad Sci 100: 11980–11985.
* Morrison D. 2012. MAPK kinase pathways. Cold Spring Harb Perspect Biol 4: a011254.
* Newton AC, Bootman MD, Scott JD. 2014. Second messengers. Cold Spring Harb Perspect Biol doi:
10.1101/cshperspect.a005926.
Novak B, Tyson JJ. 1993. Numerical analysis of a comprehensive model of M-phase control in
Xenopus oocyte extracts and intact embryos. J Cell Sci 106: 1153–1168.
Pieroni JP, Jacobowitz O, Chen J, Iyengar R. 1993. Signal recognition and integration by Gs-stimulated
adenylyl cyclases. Curr Opin Neurobiol 3: 345–351.
Pomerening JR, Kim SY, Ferrell JE Jr. 2005. Systems-level dissection of the cell-cycle oscillator:
Bypassing positive feedback produces damped oscillations. Cell 122: 565–578.
Pool IdS, Milgram S, Newcomb T. 1989. The small world. Ablex, Norwood, NJ.
* Sassone-Corsi P. 2012. The cyclic AMP pathway. Cold Spring Harb Perspect Biol 4: a011148.
* Sever R, Brugge JS. 2014. Signal transduction in cancer. Cold Spring Harb Perspect Med doi:
10.1101/cshperspect.a006098.
Sha W, Moore J, Chen K, Lassaletta AD, Yi CS, Tyson JJ, Sible JC. 2003. Hysteresis drives cell-cycle
transitions in Xenopus laevis egg extracts. Proc Natl Acad Sci 100: 975–980.
Tanaka K, Augustine GJ. 2008. A positive feedback signal transduction loop determines timing of
cerebellar long-term depression. Neuron 59: 608–620.
Tyson JJ, Novak B. 2010. Functional motifs in biochemical reaction networks. Annu Rev Phys Chem
61: 219–240.
Weng G, Bhalla US, Iyengar R. 1999. Complexity in biological signaling systems. Science 284: 92–96.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a005934
SECTION II
SIGNALING PATHWAYS
MAP Kinase Pathways
Deborah K. Morrison
Laboratory of Cell and Developmental Signaling, National Cancer Institute,
Frederick, Maryland 21702
Correspondence: morrisod@mail.nih.gov
The MAP kinases can be grouped into three main families. In mammals,
these are ERKs (extracellular-signal-regulated kinases), JNKs (Jun amino-
terminal kinases), and p38/SAPKs (stress-activated protein kinases). ERK
family members possess a TEY motif in the activation segment and can be
subdivided into two groups: the classic ERKs that consist mainly of a kinase
domain (ERK1 and ERK2) and the larger ERKs (such as ERK5) that contain
a much more extended sequence carboxy-terminal to their kinase domain
(Zhang and Dong 2007). The classic ERK1/2 module (Fig. 2) responds
primarily to growth factors and mitogens to induce cell growth and
differentiation (McKay and Morrison 2007; Shaul and Seger 2007).
Important upstream regulators of this module include cell surface receptors,
such as receptor tyrosine kinases (RTKs), G-protein-coupled receptors
(GPCRs), and integrins, as well as the small GTPases Ras and Rap.
MAPKKs for the classic ERK1/2 module are MEK1 and MEK2, and the
MAPKKKs include members of the Raf family, Mos, and Tpl2.
Figure 2. The ERK MAPK pathway.
JNK family members contain a TPY motif in the activation segment and
include JNK1, JNK2, and JNK3. The JNK module (Fig. 3) is activated by
environmental stresses (ionizing radiation, heat, oxidative stress, and DNA
damage) and inflammatory cytokines, as well as growth factors, and
signaling to the JNK module often involves the Rho family GTPases Cdc42
and Rac (Johnson and Nakamura 2007). The JNK module plays an important
role in apoptosis, inflammation, cytokine production, and metabolism
(Dhanasekaran and Reddy 2008; Huang et al. 2009; Rincon and Davis 2009).
MAPKKs for the JNK module are MKK4 and MKK7, and the MAPKKKs
include MEKK1 and MEKK4, MLK2 and MLK3, ASK1, TAK1, and Tpl2.
For all of the MAPK modules, specific scaffold proteins (Good et al.
2011) have been identified that dock at least two of the core kinases of the
module. These scaffolds contribute to MAPK signaling by increasing the
local concentration of the components, providing spatial temporal regulation
of cascade activation, and/or localizing the module to specific cellular sites or
substrates. Scaffold proteins involved in MAPK cascade signaling include
KSR and MP1 for the ERK module; JIP1, JIP2, JIP3, JIP4, and POSH for the
JNK module; and JIP2, JIP4, and OSM for the p38 module (Dhanasekaran et
al. 2007).
Figure 1 adapted, with permission, from Cell Signaling Technology (http://www.cellsignal.com).
REFERENCES
Bardwell L, Thorner J. 1996. A conserved motif at the amino termini of MEKs might mediate high-
affinity interaction with the cognate MAPKs. Trends Biochem Sci 21: 373–374.
Cuadrado A, Nebreda AR. 2010. Mechanisms and functions of p38 MAPK signalling. Biochem J 429:
403–417.
Cuenda A, Rousseau S. 2007. p38 MAP-kinases pathway regulation, function and role in human
diseases. Biochim Biophys Acta 1773: 1358–1375.
Cuevas BD, Abell AN, Johnson GL. 2007. Role of mitogen-activated protein kinase kinase kinases in
signal integration. Oncogene 26: 3159–3171.
Dhanasekaran DN, Reddy EP. 2008. JNK signaling in apoptosis. Oncogene 27: 6245–6251.
Dhanasekaran DN, Kashef K, Lee CM, Xu H, Reddy EP. 2007. Scaffold proteins of MAP-kinase
modules. Oncogene 26: 3185–3202.
Good MC, Zalatan JG, Lim WA. 2011. Scaffold proteins: Hubs for controlling the flow of cellular
information. Science 332: 680–686.
Huang G, Shi LZ, Chi H. 2009. Regulation of JNK and p38 MAPK in the immune system: Signal
integration, propagation and termination. Cytokine 48: 161–169.
Johnson GL, Nakamura K. 2007. The c-Jun kinase/stress-activated pathway: Regulation, function and
role in human disease. Biochim Biophys Acta 1773: 1341–1348.
Keshet Y, Seger R. 2010. The MAP kinase signaling cascades: A system of hundreds of components
regulates a diverse array of physiological functions. Methods Mol Biol 661: 3–38.
McKay MM, Morrison DK. 2007. Integrating signals from RTKs to ERK/MAPK. Oncogene 26: 3113–
3121.
Qi M, Elion EA. 2005. MAP kinase pathways. J Cell Sci 118: 3569–3572.
Raman M, Chen W, Cobb MH. 2007. Differential regulation and properties of MAPKs. Oncogene 26:
3100–3112.
Rincon M, Davis RJ. 2009. Regulation of the immune response by stress-activated protein kinases.
Immunol Rev 228: 212–224.
Shaul YD, Seger R. 2007. The MEK/ERK cascade: From signaling specificity to diverse functions.
Biochim Biophys Acta 1773: 1213–1226.
Tanoue T, Nishida E. 2003. Molecular recognitions in MAP kinase cascades. Cell Signal 15: 455–462.
Zhang Y, Dong C. 2007. Regulatory mechanisms of mitogen-activated kinase signaling. Cell Mol Life
Sci 64: 2771–2789.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a011254
The PI3K-PKB/Akt Pathway
Figures adapted, with kind permission, from Cell Signaling Technology (http://www.cellsignal.com).
REFERENCES
Alessi DR. 2001. Discovery of PDK1, one of the missing links in insulin signal transduction. Colworth
Medal Lecture. Biochem Soc Trans 29: 1–14.
Alessi DR, James SR, Downes CP, Holmes AB, Gaffney PR, Reese CB, Cohen P. 1997.
Characterization of a 3-phosphoinositide-dependent protein kinase which phosphorylates and
activates protein kinase Bα. Curr Biol 7: 261–269.
Altomare DA, Testa JR. 2005. Perturbations of the AKT signaling pathway in human cancer. Oncogene
24: 7455–7464.
Andjelković M, Jakubowicz T, Cron P, Ming XF, Han JW, Hemmings BA. 1996. Activation and
phosphorylation of a pleckstrin homology domain containing protein kinase (RAC-PK/PKB)
promoted by serum and protein phosphatase inhibitors. Proc Natl Acad Sci 93: 5699–5704.
Bozulic L, Hemmings BA. 2009. PIKKing on PKB: Regulation of PKB activity by phosphorylation.
Curr Opin Cell Biol 21: 256–261.
Brazil DP, Hemmings BA. 2001. Ten years of protein kinase B signalling: A hard Akt to follow.
Trends Biochem Sci 26: 657–664.
Brognard J, Sierecki E, Gao T, Newton AC. 2007. PHLPP and a second isoform, PHLPP2,
differentially attenuate the amplitude of Akt signaling by regulating distinct Akt isoforms. Mol Cell
25: 917–931.
Feng J, Park J, Cron P, Hess D, Hemmings BA. 2004. Identification of a PKB/Akt hydrophobic motif
Ser-473 kinase as DNA-dependent protein kinase. J Biol Chem 279: 41189–41196.
Guertin DA, Stevens DM, Thoreen CC, Burds AA, Kalaany NY, Moffat J, Brown M, Fitzgerald KJ,
Sabatini DM. 2006. Ablation in mice of the mTORC components raptor, rictor, or mLST8 reveals
that mTORC2 is required for signaling to Akt-FOXO and PKCα, but not S6K1. Dev Cell 11: 859–
871.
Manning BD, Cantley LC. 2007. AKT/PKB signaling: Navigating downstream. Cell 129: 1261–1274.
Sarbassov DD, Guertin DA, Ali SM, Sabatini DM. 2005. Phosphorylation and regulation of Akt/PKB
by the rictor–mTOR complex. Science 307: 1098–1101.
Stambolic V, Suzuki A, de la Pompa JL, Brothers GM, Mirtsos C, Sasaki T, Ruland J, Penninger JM,
Siderovski DP, Mak TW. 1998. Negative regulation of PKB/Akt-dependent cell survival by the
tumor suppressor PTEN. Cell 95: 29–39.
Vander Haar E, Lee SI, Bandhakavi S, Griffin TJ, Kim DH. 2007. Insulin signalling to mTOR
mediated by the Akt/PKB substrate PRAS40. Nat Cell Biol 9: 316–323.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a011189
mTOR Signaling
REFERENCES
Brown EJ, Albers MW, Shin TB, Ichikawa K, Keith CT, Lane WS, Schreiber SL. 1994. A mammalian
protein targeted by G1-arresting rapamycin-receptor complex. Nature 369: 756–758.
Heitman J, Movva NR, Hall MN. 1991. Targets for cell cycle arrest by the immunosuppressant
rapamycin in yeast. Science 253: 905–909.
Kunz J, Henriquez R, Schneider U, Deuter-Reinhard M, Movva NR, Hall MN. 1993. Target of
rapamycin in yeast, TOR2, is an essential phosphatidylinositol kinase homolog required for G1
progression. Cell 73: 585–596.
Laplante M, Sabatini DM. 2009a. An emerging role of mTOR in lipid biosynthesis. Curr Biol 19:
R1046–R1052.
Laplante M, Sabatini DM. 2009b. mTOR signaling at a glance. J Cell Sci 122: 3589–3594.
Ma XM, Blenis J. 2009. Molecular mechanisms of mTOR-mediated translational control. Nat Rev Mol
Cell Biol 10: 307–318.
Sabatini DM, Erdjument-Bromage H, Lui M, Tempst P, Snyder SH. 1994. RAFT1: A mammalian
protein that binds to FKBP12 in a rapamycin-dependent fashion and is homologous to yeast TORs.
Cell 78: 35–43.
Sabers CJ, Martin MM, Brunn GJ, Williams JM, Dumont FJ, Wiederrecht G, Abraham RT. 1995.
Isolation of a protein target of the FKBP12–rapamycin complex in mammalian cells. J Biol Chem
270: 815–822.
Vezina C, Kudelski A, Sehgal SN. 1975. Rapamycin (AY-22,989), a new antifungal antibiotic. I.
Taxonomy of the producing streptomycete and isolation of the active principle. J Antibiot (Tokyo)
28: 721–726.
Zoncu R, Efeyan A, Sabatini DM. 2011. mTOR: From growth signal integration to cancer, diabetes and
ageing. Nat Rev Mol Cell Biol 12: 21–35.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a011593
Calcium Signaling
Martin D. Bootman
The Babraham Institute Babraham Research Campus, Cambridge CB22 3AT,
United Kingdom
Correspondence: martin.bootman@bbsrc.ac.uk
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a011171
The Cyclic AMP Pathway
Paolo Sassone-Corsi
Center for Epigenetics and Metabolism, School of Medicine, University of
California, Irvine, California 92697
Correspondence: psc@uci.edu
REFERENCES
Bos JL. 2003. Epac: A new cAMP target and new avenues in cAMP research. Nat Rev Mol Cell Biol 4:
733–738.
Bruce JI, Straub SV, Yule DI. 2003. Crosstalk between cAMP and Ca2+ signaling in non-excitable
cells. Cell Calcium 34: 431–444.
Fimia GM, Sassone-Corsi P. 2001. Cyclic AMP signaling. J Cell Sci 114: 1971–1972.
Goraya TA, Cooper DMF. 2005. Ca2+-calmodulin-dependent phosphodiesterase (PDE1): Current
perspectives. Cell Signal 17: 789–797.
Mayr B, Montminy M. 2001. Transcriptional regulation by the phosphorylation-dependent factor
CREB. Nat Rev Mol Cell Biol 2: 599–609.
McKnight GS. 1991. Cyclic AMP second messenger systems. Curr Opin Cell Biol 3: 213–217.
Pierce KL, Premont RT, Lefkowitz RJ. 2002. Seven-transmembrane receptors. Nat Rev Mol Cell Biol
3: 639–650.
Sassone-Corsi P. 1995. Transcription factors responsive to cAMP. Annu Rev Cell Dev Biol 11: 355–
377.
Sutherland EW, Rall TW. 1958. Fractionation and characterization of a cyclic adenine ribonucleotide
formed by tissue particles. J Biol Chem 232: 1077–1091.
Tasken K, Skalhegg BS, Tasken KA, Solberg R, Knutsen HK, Levy FO, Sandberg M, Orstavik S,
Larsen T, Johansen AK, et al. 1997. Structure, function and regulation of human cAMP-dependent
protein kinases. Adv Second Messenger Phosphoprotein Res 31: 191–203.
Taylor SS, Knighton DR, Zheng J, Ten Eyck LF, Sowadski JM. 1992. Structural framework for the
protein kinase family. Annu Rev Cell Biol 8: 429–462.
Wong W, Scott JD. 2004. AKAP signaling complexes: Focal points in space and time. Nat Rev Mol
Cell Biol 5: 959–970.
Yoshimasa T, Sibley DR, Bouvier M, Lefkowitz RJ, Caron MG. 1987. Cross-talk between cellular
signaling pathways suggested by phorbol-ester induced adenylate cyclase phosphorylation. Nature
327: 67–70.
Zaccolo M, Pozzan T. 2003. cAMP and Ca2+ interplay: A matter of oscillation patterns. Trends
Neurosci 26: 53–55.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a011148
Wnt Signaling
Roel Nusse
Howard Hughes Medical Institute, Stanford University Medical Center,
Stanford, California 94305-5428
Correspondence: rnusse@stanford.edu
Members of the Wnt family are secreted ligands that regulate numerous
developmental pathways (Cadigan and Peifer 2009; Van Amerongen and
Nusse 2009; and see Ch. 10 [Perrimon et al. 2013]). Wnt binds to members of
the Frizzled family, activating a canonical signaling pathway that targets
members of the LEF/TCF transcription factor family (Fig. 1). These control
gene expression programs that regulate cell fate and morphogenesis (Van
Amerongen and Nusse 2009). Wnt also activates so-called non-canonical
pathways (Fig. 2), which regulate planar cell polarity by stimulating
cytoskeletal reorganization and can also lead to calcium mobilization
(Veeman et al. 2003).
Figure 1. Canonical Wnt signaling.
Figure 2. Non-canonical Wnt signaling pathways.
REFERENCES
*Reference is in this book.
Cadigan KM, Peifer M. 2009. Wnt signaling from development to disease: Insights from model
systems. Cold Spring Harb Perspect Biol 1: a002881.
Grigoryan T, Wend P, Klaus A, Birchmeier W. 2008. Deciphering the function of canonical Wnt
signals in development and disease: Conditional loss- and gain-of-function mutations of β-catenin
in mice. Genes Dev 22: 2308–2341.
* Hemmings BA, Restuccia DF. 2012. The P13K-PKB/Akt pathway. Cold Spring Harb Perspect Biol
4: a011189.
* Laplante M, Sabatini DM. 2012. mTOR signaling. Cold Spring Harb Perspect Biol 4: a011593.
* Perrimon N, Pitsouli C, Shilo B-Z. 2012. Signaling mechanisms controlling cell fate and embryonic
patterning. Cold Spring Harb Perspect Biol 4: a005975.
van Amerongen R, Nusse R. 2009. Towards an integrated view of Wnt signaling in development.
Development 136: 3205–3214.
van Amerongen R, Mikels A, Nusse R. 2008. Alternative Wnt signaling is initiated by distinct
receptors. Sci Signal 1: re9.
Veeman MT, Axelrod JD, Moon RT. 2003. A second canon. Functions and mechanisms of β-catenin-
independent Wnt signaling. Dev Cell 5: 367–377.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a011163
Hedgehog Signaling
Philip W. Ingham
Institute of Molecular and Cell Biology, Singapore 138673
Correspondence: pingham@imcb.a-star.edu.sg
REFERENCES
Ayers KL, Thérond PP. 2010. Evaluating smoothened as a G-protein-coupled receptor for Hedgehog
signalling. Trends Cell Biol 20: 287–298.
Beachy PA, Cooper MK, Young KE, von Kessler DP, Park WJ, Hall TM, Leahy DJ, Porter JA. 1997.
Multiple roles of cholesterol in hedgehog protein biogenesis and signaling. Cold Spring Harb Symp
Quant Biol 62: 191–204.
Beachy PA, Hymowitz SG, Lazarus RA, Leahy DJ, Siebold C. 2010. Interactions between Hedgehog
proteins and their binding partners come into view. Genes Dev 24: 2001–2012.
Goetz SC, Ocbina PJ, Anderson KV. 2009. The primary cilium as a Hedgehog signal transduction
machine. Methods Cell Biol 94: 199–222.
Ingham PW, Nakano Y, Seger C. 2011. Mechanisms and functions of Hedgehog signalling across the
metazoa. Nat Rev Genet 12: 393–406.
Rohatgi R, Scott MP. 2007. Patching the gaps in Hedgehog signalling. Nat Cell Biol 9: 1005–1009.
Tukachinsky H, Lopez LV, Salic A. 2010. A mechanism for vertebrate Hedgehog signaling:
Recruitment to cilia and dissociation of SuFu-Gli protein complexes. J Cell Biol 191: 415–428.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a011221
Notch Signaling
Raphael Kopan
Department of Developmental Biology, and Department of Medicine
(Division of Dermatology), Washington University, St. Louis, Missouri
63110
Correspondence: kopan@wustl.edu
The Notch pathway regulates cell proliferation, cell fate, differentiation, and
cell death in all metazoans. Notch itself is a cell-surface receptor that
transduces short-range signals by interacting with transmembrane ligands
such as Delta (termed Delta-like in humans) and Serrate (termed Jagged in
humans) on neighboring cells (Fig. 1). Some soluble ligands have also been
identified in Caenorhabditis elegans, but these bind to Notch together with
transmembrane adaptors (Komatsu et al. 2008). Ligand binding leads to
cleavage and release of the Notch intracellular domain (NICD), which then
travels to the nucleus to regulate transcriptional complexes containing the
DNA-binding protein CBF1/RBPjk/Su(H)/Lag1 (CSL).
Figure 1. Notch signaling (simplified view).
REFERENCES
*Reference is in this book.
Agrawal N, Frederick MJ, Pickering CR, Bettegowda C, Chang K, Li RJ, Fakhry C, Xie TX, Zhang J,
Wang J, et al. 2011. Exome sequencing of head and neck squamous cell carcinoma reveals
inactivating mutations in NOTCH1. Science 333: 1154–1157.
Bozkulak EC, Weinmaster G. 2009. Selective use of ADAM10 and ADAM17 in activation of Notch1
signaling. Mol Cell Biol 29: 5679–5695.
Demehri S, Liu Z, Lee J, Lin MH, Crosby SD, Roberts CJ, Grigsby PW, Miner JH, Farr AG, Kopan R.
2008. Notch-deficient skin induces a lethal systemic B-lymphoproliferative disorder by secreting
TSLP, a sentinel for epidermal integrity. PLoS Biol 6: e123.
Gordon WR, Vardar-Ulu D, L’Heureux S, Ashworth T, Malecki MJ, Sanchez-Irizarry C, McArthur
DG, Histen G, Mitchell JL, Aster JC, et al. 2009. Effects of S1 cleavage on the structure, surface
export, and signaling activity of human Notch1 and Notch2. PLoS ONE 4: e6613.
Komatsu H, Chao MY, Larkins-Ford J, Corkins ME, Somers GA, Tucey T, Dionne HM, White JQ,
Wani K, Boxem M, et al. 2008. OSM-11 facilitates LIN-12 Notch signaling during C. elegans
vulval development. PLoS Biol 6: e196.
Kopan R, Ilagan MX. 2009. The canonical Notch signaling pathway: Unfolding the activation
mechanism. Cell 137: 216–233.
Logeat F, Bessia C, Brou C, Lebail O, Jarriault S, Seidah NG, Israël A. 1998. The Notch1 receptor is
cleaved constitutively by a furin-like convertase. Proc Natl Acad Sci 95: 8108–8112.
Luty WH, Rodeberg D, Parness J, Vyas YM. 2007. Antiparallel segregation of Notch components in
the immunological synapse directs reciprocal signaling in allogeneic Th:DC conjugates. J Immunol
179: 819–829.
Moberg KH, Schelble S, Burdick SK, Hariharan IK. 2005. Mutations in erupted, the Drosophila
ortholog of mammalian tumor susceptibility gene 101, elicit non-cell-autonomous overgrowth. Dev
Cell 9: 699–710.
Mukherjee T, Kim WS, Mandal L, Banerjee U. 2011. Interaction between Notch and Hif-α in
development and survival of Drosophila blood cells. Science 332: 1210–1213.
* Nusse R. 2012. Wnt signaling. Cold Spring Harb Perspect Biol 4: a011163.
Ramain P, Khechumian K, Seugnet L, Arbogast N, Ackermann C, Heitzler P. 2001. Novel Notch
alleles reveal a Deltex-dependent pathway repressing neural fate. Curr Biol 11: 1729–1738.
Rangarajan A, Talora C, Okuyama R, Nicolas M, Mammucari C, Oh H, Aster JC, Krishna S, Metzger
D, Chambon P, et al. 2001. Notch signaling is a direct determinant of keratinocyte growth arrest
and entry into differentiation. EMBO J 20: 3427–3436.
Sanders PG, Munoz-Descalzo S, Balayo T, Wirtz-Peitz F, Hayward P, Arias AM. 2009. Ligand-
independent traffic of Notch buffers activated armadillo in Drosophila. PLoS Biol 7: e1000169.
Sprinzak D, Lakhanpal A, LeBon L, Santat LA, Fontes ME, Anderson GA, Garcia-Ojalvo J, Elowitz
MB. 2010. Cis-interactions between Notch and Delta generate mutually exclusive signaling states.
Nature 465: 86–90.
Stransky N, Egloff AM, Tward AD, Kostic AD, Cibulskis K, Sivachenko A, Kryukov GV, Lawrence
MS, Sougnez C, McKenna A, et al. 2011. The mutational landscape of head and neck squamous
cell carcinoma. Science 333: 1157–1160.
Tagami S, Okochi M, Yanagida K, Ikuta A, Fukumori A, Matsumoto N, Ishizuka-Katsura Y,
Nakayama T, Itoh N, Jiang J, et al. 2008. Regulation of Notch signaling by dynamic changes in the
precision in S3 cleavage of Notch-1. Mol Cell Biol 28: 165–176.
Thompson BJ, Mathieu J, Sung HH, Loeser E, Rorth P, Cohen SM. 2005. Tumor suppressor properties
of the ESCRT-II complex component Vps25 in Drosophila. Dev Cell 9: 711–720.
Vaccari T, Bilder D. 2005. The Drosophila tumor suppressor vps25 prevents nonautonomous
overproliferation by regulating Notch trafficking. Dev Cell 9: 687–698.
van Tetering G, van Diest P, Verlaan I, van der Wall E, Kopan R, Vooijs M. 2009. Metalloprotease
ADAM10 is required for Notch1 site 2 cleavage. J Biol Chem 284: 31018–31027.
Weng AP, Ferrando AA, Lee W, Morris JPIV, Silverman LB, Sanchez-Irizarry C, Blacklow SC, Look
AT, Aster JC. 2004. Activating mutations of NOTCH1 in human T cell acute lymphoblastic
leukemia. Science 306: 269–271.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a011213
Signaling by the TGFβ Superfamily
Jeffrey L. Wrana
Samuel Lunenfeld Research Institute, Mount Sinai Hospital, Toronto, Ontario
M5G 1X5, Canada
Correspondence: wrana@lunenfeld.ca
REFERENCES
Attisano L, Wrana JL. 2002. Signal transduction by the TGF-β superfamily. Science 296: 1646–1647.
Attisano L, Wrana JL. 2013. Signal integration in TGF-β, WNT and Hippo pathways. F1000Prime Rep
5: 17.
Green JB, Smith JC. 1990. Graded changes in dose of a Xenopus activin A homologue elicit stepwise
transitions in embryonic cell fate. Nature 347: 391–394.
Griswold-Prenner I, Kamibayashi C, Maruoka EM, Mumby MC, Derynck R. 1998. Physical and
functional interactions between type I transforming growth factor β receptors and Bα, a WD-40
repeat subunit of phosphatase 2A. Mol Cell Biol 18: 6595–6604.
Lee MK, Pardoux C, Hall MC, Lee PS, Warburton D, Qing J, Smith SM, Derynck R. 2007. TGF-β
activates Erk MAP kinase signalling through direct phosphorylation of ShcA. EMBO J 26: 3957–
3967.
Massague J. 1998. TGF-β signal transduction. Annu Rev Biochem 67: 753–791.
Massague J. 2012. TGFβ signalling in context. Nat Rev Mol Cell Biol 13: 616–630.
Miyazono K, Kamiya Y, Morikawa M. 2010. Bone morphogenetic protein receptors and signal
transduction. J Biochem 147: 35–51.
Morrell NW. 2011. Role of bone morphogenetic protein receptors in the development of pulmonary
arterial hypertension. Adv Exp Med Biol 661: 251–264.
Mu Y, Gudey SK, Landström M. 2012. Non-Smad signaling pathways. Cell Tissue Res 347: 11–20.
Niederlander C, Walsh JJ, Episkopou V, Jones CM. 2001. Arkadia enhances nodal-related signalling to
induce mesendoderm. Nature 410: 830–834.
Ozdamar B, Bose R, Barrios-Rodiles M, Wang HR, Zhang Y, Wrana JL. 2005. Regulation of the
polarity protein Par6 by TGF-β receptors controls epithelial cell plasticity. Science 307: 1603–1609.
Podkowa M, Zhao X, Chow CW, Coffey ET, Davis RJ, Attisano L. 2010. Microtubule stabilization by
bone morphogenetic protein receptor-mediated scaffolding of c-Jun N-terminal kinase promotes
dendrite formation. Mol Cell Biol 30: 2241–2250.
Roberts AB, Anzano MA, Lamb LC, Smith JM, Sporn MB. 1981. New class of transforming growth
factors potentiated by epidermal growth factor: Isolation from non-neoplastic tissues. Proc Natl
Acad Sci 78: 5339–5343.
Stroschein SL, Wang W, Zhou S, Zhou Q, Luo K. 1999. Negative feedback regulation of TGF-β
signaling by the SnoN oncoprotein. Science 286: 771–774.
Wilkes MC, Repellin CE, Hong M, Bracamonte M, Penheiter SG, Borg JP, Leof EB. 2009. Erbin and
the NF2 tumor suppressor Merlin cooperatively regulate cell-type-specific activation of PAK2 by
TGF-β. Dev Cell 16: 433–444.
Yi JJ, Barnes AP, Hand R, Polleux F, Ehlers MD. 2010. TGF-β signaling specifies axons during brain
development. Cell 142: 144–157.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a011197
The JAK/STAT Pathway
Douglas A. Harrison
Department of Biology, University of Kentucky, Lexington, Kentucky 40506
Correspondence: DougH@uky.edu
In mammals, there are four members of the JAK family and seven
STATs. Different JAKs and STATs are recruited based on their tissue
specificity and the receptors engaged in the signaling event (Schindler and
Plumlee 2008). In invertebrates, the Drosophila JAK/STAT pathway has
been extensively studied and comprises only one JAK and one STAT
(Arbouzova and Zeidler 2006). Although the canonical JAK/STAT pathway
is simple and direct, pathway components regulate or are regulated by
members of other signaling pathways, including those involving the ERK
MAP kinase, PI 3-kinase (PI3K), and others. Furthermore, non-canonical
JAK and STAT activities influence the global transcriptional state through
modification of chromatin structure (Li 2008; Dawson et al. 2009).
Human JAK mutations cause numerous diseases, including severe
combined immune deficiency, hyperIgE syndrome, certain leukemias,
polycythemia vera, and other myeloproliferative disorders (Jatiani et al.
2010). Because of the causative role in these diseases and their central
significance in immune response, JAKs have become attractive targets for
development of therapeutics for a variety of hematopoietic and immune
system disorders (Pesu et al. 2008; Haan et al. 2010). Owing to the pleiotropy
of the JAK/STAT pathway, agents that selectively perturb specific family
members are being sought.
Figures adapted by kind permission of Cell Signaling Technology (http://cellsignal.com).
REFERENCES
Arbouzova NI, Zeidler MP. 2006. JAK/STAT signalling in Drosophila: Insights into conserved
regulatory and cellular functions. Development 133: 2605–2616.
Darnell JE Jr, Kerr IM, Stark GR. 1994. Jak-STAT pathways and transcriptional activation in response
to IFNs and other extracellular signaling proteins. Science 264: 1415–1421.
Dawson MA, Bannister AJ, Gottgens B, Foster SD, Bartke T, Green AR, Kouzarides T. 2009. JAK2
phosphorylates histone H3Y41 and excludes HP1a from chromatin. Nature 461: 819–822.
Ghoreschi K, Laurence A, O’Shea JJ. 2009. Janus kinases in immune cell signaling. Immunol Rev 228:
273–287.
Haan C, Behrmann I, Haan S. 2010. Perspectives for the use of structural information and chemical
genetics to develop inhibitors of Janus kinases. J Cell Mol Med 14: 504–527.
Jatiani SS, Baker SJ, Silverman LR, Reddy EP. 2010. JAK/STAT pathways in cytokine signaling and
myeloproliferative disorders: Approaches for targeted therapies. Genes Cancer 1: 979–993.
Li WX. 2008. Canonical and non-canonical JAK-STAT signaling. Trends Cell Biol 18: 545–551.
Pesu M, Laurence A, Kishore N, Zwillich SH, Chan G, O’Shea JJ. 2008. Therapeutic targeting of Janus
kinases. Immunol Rev 223: 132–142.
Schindler C, Plumlee C. 2008. Inteferons pen the JAK-STAT pathway. Semin Cell Dev Biol 19: 311–
318.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a011205
Toll-Like Receptor Signaling
REFERENCES
*Reference is in this book.
Barton GM, Kagan JC. 2009. A cell biological view of Toll-like receptor function: Regulation through
compartmentalization. Nat Rev Immunol 9: 535–542.
Blasius AL, Beutler B. 2010. Intracellular Toll-like receptors. Immunity 32: 305–315.
Casanova JL, Abel L, Quintana-Murci L. 2011. Human TLRs and IL-1Rs in host defense: Natural
insights from evolutionary, epidemiological, and clinical genetics. Annu Rev Immunol 29: 447–491.
Flannery S, Bowie AG. 2010. The interleukin-1 receptor-associated kinases: Critical regulators of
innate immune signalling. Biochem Pharmacol 80: 1981–1991.
Green NM, Marshak-Rothstein A. 2011. Toll-like receptor driven B cell activation in the induction of
systemic autoimmunity. Semin Immunol 23: 106–112.
Hacker H, Tseng PH, Karin M. 2011. Expanding TRAF function: TRAF3 as a tri-faced immune
regulator. Nat Rev Immunol 11: 457–468.
Hennessy EJ, Parker AE, O’Neill LA. 2010. Targeting Toll-like receptors: Emerging therapeutics? Nat
Rev Drug Discov 9: 293–307.
Kawai T, Akira S. 2010. The role of pattern-recognition receptors in innate immunity: Update on Toll-
like receptors. Nat Immunol 11: 373–384.
Lemaitre B, Nicolas E, Michaut L, Reichhart JM, Hoffmann JA. 1996. The dorsoventral regulatory
gene cassette spatzle/Toll/cactus controls the potent antifungal response in Drosophila adults. Cell
86: 973–983.
McGettrick AF, O’Neill LA. 2010. Localisation and trafficking of Toll-like receptors: An important
mode of regulation. Curr Opin Immunol 22: 20–27.
Medzhitov R, Preston-Hurlburt P, Janeway CA Jr. 1997. A human homologue of the Drosophila Toll
protein signals activation of adaptive immunity. Nature 388: 394–397.
* Morrison DK. 2012. MAP kinase pathways. Cold Spring Harb Perspect Biol 4: a011254.
* Newton K, Dixit VM. 2012. Signaling in innate immunity and inflammation. Cold Spring Harb
Perspect Biol 4: a006049.
Ngo VN, Young RM, Schmitz R, Jhavar S, Xiao W, Lim KH, Kohlhammer H, Xu W, Yang Y, Zhao
H, et al. 2011. Oncogenically active MYD88 mutations in human lymphoma. Nature 470: 115–119.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a011247
Immunoreceptor Signaling
Lawrence E. Samelson
Center for Cancer Research, National Cancer Institute, National Institutes of
Health, Bethesda, Maryland 20892-4256
Correspondence: lsamelson@comcast.net
T cells and B cells are stimulated when antigens bind to T-cell receptors
(TCRs) and B-cell receptors (BCRs) in their respective plasma membranes
(Fig. 1). The antigens presented to T cells are in the form of short peptides
that have been processed in infected cells and are displayed on the surface
bound to major histocompatibility complex (MHC) class I or class II
molecules. BCRs, in contrast, recognize free antigens in their native form
present in the extracellular milieu or on cell surfaces.
Figure 1. Early events in T cell and B cell receptor signaling.
Despite superficial differences, the responses in the two cell types are
remarkably similar. For both, signaling begins with engagement of a complex
receptor composed of antigen receptor subunits with immunoglobulin or
immunoglobulin-like domains. Both receptors also contain non-polymorphic
subunits, the TCRζ chain and CD3 subunits (γ, δ, and ε) for the TCR, and the
α and β chains of the BCR (Reth 1995; Wucherpfennig et al. 2010).
Coreceptors are important in both T cells (CD4 for helper T cells; CD8 for
cytotoxic T cells) and B cells (CD19). For both receptors, members of two
protein tyrosine kinase (PTK) families, Src (Lck for T cells and Lyn for B
cells) and Syk (Zap-70 for T cells and Syk for B cells) (Bradshaw 2010; Liu
et al. 2010; Wang et al. 2010), are critical signaling molecules closely
associated with the receptors and coreceptors. Activation of tyrosine kinases
is the first biochemical change that follows receptor engagement.
Following antigen engagement, activated tyrosine kinases phosphorylate
several adapter proteins and signaling enzymes, which then interact. For T
cells, this is exemplified by the phosphorylation of the adapters LAT and
SLP-76 (Fig. 2), the formation of multiprotein complexes nucleated at these
adapters, and the inclusion of such enzymes as phospholipase Cγ1 (PLCγ1)
and VAV in the protein assemblies (Balagopalan et al. 2010; Jordan and
Koretzky 2010). Activation of an additional protein tyrosine kinase, Itk in T
cells and Btk in B cells, occurs in these complexes (Andreotti et al. 2010).
Similarly in B-cells, phosphorylation of BLNK and CD19 leads to
recruitment of adapters and signaling enzymes. In both cell types these events
occur at the plasma membrane, where critical lipid substrates of enzymes
such as PLCγ1 and PI3 kinase (PI3K) are located.
Figure 2. T cell receptor signaling.
REFERENCES
Andreotti AH, Schwartzberg PL, Joseph RE, Berg LJ. 2010. T-cell signaling regulated by the Tec
family kinase, Itk. Cold Spring Harb Perspect Biol 2: a002287.
Balagopalan L, Coussens NP, Sherman E, Samelson LE, Sommers CL. 2010. The LAT story: A tale of
cooperativity, coordination, and choreography. Cold Spring Harb Perspect Biol 2: a005512.
Bradshaw JM. 2010. The Src, Syk, and Tec family kinases: Distinct types of molecular switches. Cell
Signal 22: 1175–1184.
Finlay D, Cantrell D. 2010. The coordination of T-cell function by serine/threonine kinases. Cold
Spring Harb Perspect Biol 3: a002261.
Huang YH, Sauer K. 2010. Lipid signaling in T-cell development and function. Cold Spring Harb
Perspect Biol 2: a002428.
Jordan MS, Koretzky GA. 2010. Coordination of receptor signaling in multiple hematopoietic cell
lineages by the adaptor protein SLP-76. Cold Spring Harb Perspect Biol 2: a002501.
Liu W, Sohn HW, Tolar P, Pierce SK. 2010. It’s all about change: The antigen-driven initiation of B-
cell receptor signaling. Cold Spring Harb Perspect Biol 2: a002295.
Reth M. 1995. The B-cell antigen receptor complex and co-receptors. Immunol Today 16: 310–313.
Thome M, Charton JE, Pelzer C, Hailfinger S. 2010. Antigen receptor signaling to NF-κB via
CARMA1, BCL10, and MALT1. Cold Spring Harb Perspect Biol 2: a003044.
Wang H, Kadlecek TA, Au-Yeung BB, Goodfellow HE, Hsu LY, Freedman TS, Weiss A. 2010. ZAP-
70: An essential kinase in T-cell signaling. Cold Spring Harb Perspect Biol 2: a002279.
Wucherpfennig KW, Gagnon E, Call MJ, Huseby ES, Call ME. 2010. Structural biology of the T-cell
receptor: Insights into receptor assembly, ligand recognition, and initiation of signaling. Cold
Spring Harb Perspect Biol 2: a005140.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a011510
Signaling by Nuclear Receptors
Correspondence: ckg@ucsd.edu
Studies of the salivary glands of insect larva in the 1960s first indicated
that steroid hormones regulate transcription. Subsequent work showed that
estrogen can selectively activate the genes encoding egg-white and yolk
proteins, leading to the cloning of the estrogen, glucocorticoid, and thyroid
hormone receptors in the 1980s (Hollenberg et al. 1985; Green et al. 1986;
Miesfeld et al. 1986; Sap et al. 1986; Weinberger et al. 1986). We now know
that 48 nuclear receptors are encoded in the human genome (Mangelsdorf et
al. 1995). In many cases, ligands for these have been identified, but several
“orphan receptors” remain (Burris et al. 2012). Whether all of these have
bona fide ligands is unclear, because some nuclear receptors can act in the
absence of a ligand (Table 1).
Table 1. Common nuclear receptors and their ligands
Receptor Abbreviation Ligand
Androgen receptor AR Testosterone
Estrogen receptor ER Estrogen
Estrogen-related receptor ERR ?
Glucocorticoid receptor GR Cortisol
Mineralocorticoid receptor MR Aldosterone
Progesterone receptor PR Progesterone
Retinoic acid receptor RAR Retinoic acid
Retinoid orphan receptor ROR ?
Retinoic acid-related receptor RXR Rexinoids
Liver X receptor LXR Oxysterols
Peroxisome proliferator-activated receptor γ PPARγ Fatty acid metabolites
Thyroid hormone receptor TR Thyroid hormone
Vitamin D3 receptor VDR Vitamin D3
REFERENCES
Berrabah W, Aumercier P, Lefebvre P, Staels B. 2011. Control of nuclear receptor activities in
metabolism by post-translational modifications. FEBS Lett 585: 1640–1650.
Boergesen M, Pedersen TA, Gross B, van Heeringen SJ, Hagenbeek D, Bindesboll C, Caron S,
Lalloyer F, Steffensen KR, Nebb HI, et al. 2012. Genome-wide profiling of liver X receptor,
retinoid X receptor, and peroxisome proliferator-activated receptor α in mouse liver reveals
extensive sharing of binding sites. Mol Cell Biol 32: 852–867.
Bulynko YA, O’Malley BW. 2011. Nuclear receptor coactivators: Structural and functional
biochemistry. Biochemistry 50: 313–328.
Burris TP, Busby SA, Griffin PR. 2012. Targeting orphan nuclear receptors for treatment of metabolic
diseases and autoimmunity. Chem Biol 19: 51–59.
Carroll JS, Meyer CA, Song J, Li W, Geistlinger TR, Eeckhoute J, Brodsky AS, Keeton EK, Fertuck
KC, Hall GF, et al. 2006. Genome-wide analysis of estrogen receptor binding sites. Nat Genet 38:
1289–1297.
Chen JD, Evans RM. 1995. A transcriptional co-repressor that interacts with nuclear hormone
receptors. Nature 377: 454–457.
Echeverria PC, Picard D. 2010. Molecular chaperones, essential partners of steroid hormone receptors
for activity and mobility. Biochim Biophys Acta 1803: 641–649.
Glass CK, Rosenfeld MG. 2000. The coregulator exchange in transcriptional functions of nuclear
receptors. Genes Dev 14: 121–141.
Green S, Walter P, Kumar V, Krust A, Bornert JM, Argos P, Chambon P. 1986. Human oestrogen
receptor cDNA: Sequence, expression and homology to v-erb-A. Nature 320: 134–139.
Heinz S, Benner C, Spann N, Bertolino E, Lin YC, Laslo P, Cheng JX, Murre C, Singh H, Glass CK.
2010. Simple combinations of lineage-determining transcription factors prime cis-regulatory
elements required for macrophage and B cell identities. Mol Cell 38: 576–589.
Hollenberg SM, Weinberger C, Ong ES, Cerelli G, Oro A, Lebo R, Thompson EB, Rosenfeld MG,
Evans RM. 1985. Primary structure and expression of a functional human glucocorticoid receptor
cDNA. Nature 318: 635–641.
Horlein AJ, Naar AM, Heinzel T, Torchia J, Gloss B, Kurokawa R, Ryan A, Kamei Y, Soderstrom M,
Glass CK, et al. 1995. Ligand-independent repression by the thyroid hormone receptor mediated by
a nuclear receptor co-repressor. Nature 377: 397–404.
Lee JH, Lee MJ. 2012. Emerging roles of the ubiquitin–proteasome system in the steroid receptor
signaling. Arch Pharm Res 35: 397–407.
Mangelsdorf DJ, Thummel C, Beato M, Herrlich P, Schutz G, Umesono K, Blumberg B, Kastner P,
Mark M, Chambon P, et al. 1995. The nuclear receptor superfamily: The second decade. Cell 83:
835–839.
Miesfeld R, Rusconi S, Godowski PJ, Maler BA, Okret S, Wikstrom AC, Gustafsson JA, Yamamoto
KR. 1986. Genetic complementation of a glucocorticoid receptor deficiency by expression of
cloned receptor cDNA. Cell 46: 389–399.
Nolte RT, Wisely GB, Westin S, Cobb JE, Lambert MH, Kurokawa R, Rosenfeld MG, Willson TM,
Glass CK, Milburn MV. 1998. Ligand binding and co-activator assembly of the peroxisome
proliferator-activated receptor-γ. Nature 395: 137–143.
Sap J, Munoz A, Damm K, Goldberg Y, Ghysdael J, Leutz A, Beug H, Vennstrom B. 1986. The c-erb-
A protein is a high-affinity receptor for thyroid hormone. Nature 324: 635–640.
Treuter E, Venteclef N. 2011. Transcriptional control of metabolic and inflammatory pathways by
nuclear receptor SUMOylation. Biochim Biophys Acta 1812: 909–918.
Watson PJ, Fairall L, Schwabe JW. 2012. Nuclear hormone receptor co-repressors: Structure and
function. Mol Cell Endocrinol 348: 440–449.
Weinberger C, Thompson CC, Ong ES, Lebo R, Gruol DJ, Evans RM. 1986. The c-erb-A gene encodes
a thyroid hormone receptor. Nature 324: 641–646.
Wu Q, Chambliss K, Umetani M, Mineo C, Shaul PW. 2011. Non-nuclear estrogen receptor signaling
in the endothelium. J Biol Chem 286: 14737–14743.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a016709
The Hippo Pathway
Correspondence: ikh@berkeley.edu
Core
Hippo (HPO) MST1 and MST2
Salvador (SAV) SAV1
Mats (MTS) MOBKL1A and MOBKL1AB
Warts (WTS) LATS1 and LATS2
Downstream
Yorkie (YKI) YAP and TAZ
WBP2 WBP2
Scalloped (SD) TEAD1—TEAD4
MAD SMADs
TSH TSHZ1—TSHZ3
HTH MEIS1—MEIS3
REFERENCES
Boggiano JC, Vanderzalm PJ, Fehon RG. 2011. Tao-1 phosphorylates Hippo/MST kinases to regulate
the Hippo-Salvador-Warts tumor suppressor pathway. Dev Cell 21: 888–895.
Gilbert MM, Tipping M, Veraksa A, Moberg KH. 2011. A screen for conditional growth suppressor
genes identifies the Drosophila homolog of HD-PTP as a regulator of the oncoprotein Yorkie. Dev
Cell 20: 700–712.
Grusche FA, Richardson HE, Harvey KF. 2010. Upstream regulation of the hippo size control pathway.
Curr Biol 20: R574–582.
Halder G, Johnson RL. 2011. Hippo signaling: Growth control and beyond. Development 138: 9–22.
Hamaratoglu F, Willecke M, Kango-Singh M, Nolo R, Hyun E, Tao C, Jafar-Nejad H, Halder G. 2006.
The tumour-suppressor genes NF2/Merlin and Expanded act through Hippo signalling to regulate
cell proliferation and apoptosis. Nat Cell Biol 8: 27–36.
Harvey K, Tapon N. 2007. The Salvador-Warts-Hippo pathway—an emerging tumour-suppressor
network. Nat Rev Cancer 7: 182–191.
Harvey KF, Pfleger CM, Hariharan IK. 2003. The Drosophila Mst ortholog, hippo, restricts growth and
cell proliferation and promotes apoptosis. Cell 114: 457–467.
Huang J, Wu S, Barrera J, Matthews K, Pan D. 2005. The Hippo signaling pathway coordinately
regulates cell proliferation and apoptosis by inactivating Yorkie, the Drosophila homolog of YAP.
Cell 122: 421–434.
Jia J, Zhang W, Wang B, Trinko R, Jiang J. 2003. The Drosophila Ste20 family kinase dMST functions
as a tumor suppressor by restricting cell proliferation and promoting apoptosis. Genes Dev 17:
2514–2519.
Justice RW, Zilian O, Woods DF, Noll M, Bryant PJ. 1995. The Drosophila tumor suppressor gene
warts encodes a homolog of human myotonic dystrophy kinase and is required for the control of
cell shape and proliferation. Genes Dev 9: 534–546.
Kango-Singh M, Nolo R, Tao C, Verstreken P, Hiesinger PR, Bellen HJ, Halder G. 2002. Shar-pei
mediates cell proliferation arrest during imaginal disc growth in Drosophila. Development 129:
5719–5730.
Lai ZC, Wei X, Shimizu T, Ramos E, Rohrbaugh M, Nikolaidis N, Ho LL, Li Y. 2005. Control of cell
proliferation and apoptosis by mob as tumor suppressor, mats. Cell 120: 675–685.
Oh H, Irvine KD. 2010. Yorkie: The final destination of Hippo signaling. Trends Cell Biol 20: 410–
417.
Pan D. 2010. The hippo signaling pathway in development and cancer. Dev Cell 19: 491–505.
Pantalacci S, Tapon N, Leopold P. 2003. The Salvador partner Hippo promotes apoptosis and cell-cycle
exit in Drosophila. Nat Cell Biol 5: 921–927.
Poon CL, Lin JI, Zhang X, Harvey KF. 2011. The sterile 20-like kinase Tao-1 controls tissue growth by
regulating the Salvador-Warts-Hippo pathway. Dev Cell 21: 896–906.
Rauskolb C, Pan G, Reddy BV, Oh H, Irvine KD. 2011. Zyxin links fat signaling to the hippo pathway.
PLoS Biol 9: e1000624.
Ribeiro PS, Josue F, Wepf A, Wehr MC, Rinner O, Kelly G, Tapon N, Gstaiger M. 2010. Combined
functional genomic and proteomic approaches identify a PP2A complex as a negative regulator of
Hippo signaling. Mol Cell 39: 521–534.
Schlegelmilch K, Mohseni M, Kirak O, Pruszak J, Rodriguez JR, Zhou D, Kreger BT, Vasioukhin V,
Avruch J, Brummelkamp TR, et al. 2011. Yap1 acts downstream of alpha-catenin to control
epidermal proliferation. Cell 144: 782–795.
Tapon N, Harvey KF, Bell DW, Wahrer DC, Schiripo TA, Haber DA, Hariharan IK. 2002. salvador
promotes both cell cycle exit and apoptosis in Drosophila and is mutated in human cancer cell
lines. Cell 110: 467–478.
Udan RS, Kango-Singh M, Nolo R, Tao C, Halder G. 2003. Hippo promotes proliferation arrest and
apoptosis in the Salvador/Warts pathway. Nat Cell Biol 5: 914–920.
Wu S, Huang J, Dong J, Pan D. 2003. hippo encodes a Ste-20 family protein kinase that restricts cell
proliferation and promotes apoptosis in conjunction with salvador and warts. Cell 114: 445–456.
Xu T, Wang W, Zhang S, Stewart RA, Yu W. 1995. Identifying tumor suppressors in genetic mosaics:
The Drosophila lats gene encodes a putative protein kinase. Development 121: 1053–1063.
Zhang X, Milton CC, Poon CL, Hong W, Harvey KF. 2011. Wbp2 cooperates with Yorkie to drive
tissue growth downstream of the Salvador-Warts-Hippo pathway. Cell Death Differ 18: 1346–
1355.
Zhao B, Tumaneng K, Guan KL. 2011. The Hippo pathway in organ size control, tissue regeneration
and stem cell self-renewal. Nat Cell Biol 13: 877–883.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a011288
SECTION III
SIGNALING PROCESSES
CHAPTER 5
SUMMARY
Outline
1 Introduction
2 Transcriptional regulation of G1 cyclins by mitogenic signals
3 Transcriptional regulation of CDK inhibitors
4 Control of G1 cyclins by the ubiquitin–proteasome system
5 Control of G1 CDK inhibitors by the ubiquitin–proteasome system
6 Concluding remarks
References
1 INTRODUCTION
Control of cell proliferation generally occurs during the first gap phase (G1)
of the eukaryotic cell division cycle (see Box 1). Multiple signals, ranging
from growth factors to DNA damage to developmental cues, influence the
decision to enter S phase, when DNA is replicated (Fig. 1). Hence, G1 phase
cell cycle control is intrinsically linked with a diverse set of pathways
controlling differentiation, stem and progenitor cell quiescence, senescence,
and responses to a variety of stresses. The decision to enter S phase from G1
represents a point of no return that, in the absence of stress such as DNA
damage, commits cells to complete the cell cycle and divide, and is therefore
tightly controlled. This decision is made at what is called the “restriction
point” in mammalian cells and “START” in yeast, after which cells become
largely refractory to extracellular signals and will complete S phase and
proceed through a second gap phase (G2 phase) and then mitosis. In
multicellular organisms, most differentiated cells exit the active cell cycle
during G1 phase and enter G0 phase, in which they remain metabolically
active for days or even years, performing specialized functions. Postmitotic
nerve and skeletal muscle cells provide good examples. Some G0 cells, such
as quiescent T cells, can be stimulated by mitogenic signals to reenter the cell
cycle.
Figure 1. G1 cell cycle control by the pRB pathway. Many cellular signaling events are intrinsically
linked to G1 phase of the cell cycle, which is controlled by the RB pathway. Signaling to the RB
pathway, and thus G1 control by different cellular processes, is achieved mainly through the regulation
of cyclins and CDK inhibitors (CKIs). In mammalian cells, mitogenic signals first induce the synthesis
of D-type cyclins, leading to activation of cyclin-D-dependent CDK4 and CDK6, and then induce E-
type cyclins to activate CDK2. Cyclin-D–CDK4/6 and cyclin-E–CDK2 cooperatively phosphorylate
RB-family proteins, derepressing E2F to allow transcription of E2F-target genes, thereby promoting the
G1/S transition. The INK4 proteins specifically inhibit CDK4 and CDK6, whereas the p21 (CIP/KIP)
family of CKIs inhibits multiple CDKs. Although the schematic illustration is based on mammalian
cells, the regulation of both G1 cyclins and CDK inhibitors is evolutionarily conserved.
The classical cell cycle comprises four phases—G1, S, G2, and M—and
is controlled by cyclin-dependent kinases (CDKs) and their cyclin
partners. The commitment to divide occurs in G1 phase, which is
controlled by cyclin-D–CDK4/6 and cyclin-E–CDK2 at the so-called
G1/S transition. DNA is then replicated in S phase. This is followed by a
second gap phase, G2, at the end of which cyclin-B–CDK1 controls
entry into M phase (mitosis), when the cell divides. Cells can exit the
cell cycle in G1 phase and enter G0 phase (quiescence). In some cases,
they can reenter the cell cycle and begin dividing again (see main text).
Figure 4. Targeting ubiquitin-dependent degradation of CDK inhibitors. The p21 family of CKIs is
regulated by the ubiquitin–proteasome pathway. In many cases, this involves phosphorylation of these
CKIs. Phosphorylated CKIs are recognized by F-box proteins such as Cdc4 in budding yeast or SKP2
in human cells, which, through the SKP1 linker protein, recruits the CKI substrate to the SCF E3 ligase
for ubiquitylation.
6 CONCLUDING REMARKS
Precise cell cycle regulation is an essential aspect of normal development and
adult homeostasis. To achieve this, cells in G1 phase integrate inputs from
major cellular signaling pathways to decide whether or not to enter S phase,
which is an irreversible cell cycle step. This integration of signals is
transformed into an appropriate level of CDK activity in large part via
changes in the level of cyclins and CKIs achieved through the regulation of
both transcription and protein stability. One challenge for the future is to
understand how multiple signaling pathways cooperate to precisely regulate
cyclin and CKI activity in various cell types, particularly stem cells, in intact
tissues. Another is to use this information to develop novel therapeutics for
the treatment of cancer, which arises in part because of disruptions to
signaling pathways that affect cell cycle regulation.
ACKNOWLEDGMENTS
We thank Alan Diehl, Andrew Koff, Kun-Liang Guan, Michele Pagano, DJ
Pan, and Xin-Hai Pei for discussions, and Tadashi Nakagawa and Ruiting Lin
for helping with figure preparation.
REFERENCES
*Reference is in this book.
Abbas T, Sivaprasad U, Terai K, Amador V, Pagano M, Dutta A. 2008. PCNA-dependent regulation of
p21 ubiquitylation and degradation via the CRL4Cdt2 ubiquitin ligase complex. Genes Dev 22:
2496–2506.
Akiyama T, Ohuchi T, Sumida S, Matsumoto K, Toyoshima K. 1992. Phosphorylation of the
retinoblastoma protein by cdk2. Proc Natl Acad Sci 89: 7900–7904.
Albanese C, Johnson J, Watanabe G, Eklund N, Vu D, Arnold A, Pestell RG. 1995. Transforming
p21ras mutants and c-Ets-2 activate the cyclin D1 promoter through distinguishable regions. J Biol
Chem 270: 23589–23597.
Bates S, Peters G. 1995. Cyclin D1 as a cellular proto-oncogene. Semin Cancer Biol 6: 73–82.
Baugh LR, Sternberg PW. 2006. DAF-16/FOXO regulates transcription of cki-1/Cip/Kip and
repression of lin-4 during C. elegans L1 arrest. Curr Biol 16: 780–785.
Bendjennat M, Boulaire J, Jascur T, Brickner H, Barbier V, Sarasin A, Fotedar A, Fotedar R. 2003. UV
irradiation triggers ubiquitin-dependent degradation of p21WAF1 to promote DNA repair. Cell
114: 599–610.
Berthet C, Aleem E, Coppola V, Tessarollo L, Kaldis P. 2003. Cdk2 knockout mice are viable. Curr
Biol 13: 1775–1785.
Bornstein G, Bloom J, Sitry-Shevah D, Nakayama K, Pagano M, Hershko A. 2003. Role of the
SCFSkp2 ubiquitin ligase in the degradation of p21Cip1 in S phase. J Biol Chem 278: 25752–
25757.
Bracken AP, Kleine-Kohlbrecher D, Dietrich N, Pasini D, Gargiulo G, Beekman C, Theilgaard-Monch
K, Minucci S, Porse BT, Marine JC, et al. 2007. The Polycomb group proteins bind throughout the
INK4A-ARF locus and are disassociated in senescent cells. Genes Dev 21: 525–530.
Brugarolas J, Chandrasekaran C, Gordon JI, Beach D, Jacks T, Hannon GJ. 1995. Radiation-induced
cell cycle arrest compromised by p21 deficiency. Nature 377: 552–557.
Buttitta LA, Katzaroff AJ, Perez CL, de la Cruz A, Edgar BA. 2007. A double-assurance mechanism
controls cell cycle exit upon terminal differentiation in Drosophila. Dev Cell 12: 631–643.
Cao X, Pfaff SL, Gage FH. 2008. YAP regulates neural progenitor cell number via the TEA domain
transcription factor. Genes Dev 22: 3320–3334.
Clayton JE, van den Heuvel SJ, Saito RM. 2008. Transcriptional control of cell-cycle quiescence
during C. elegans development. Dev Biol 313: 603–613.
Correa-Bordes J, Nurse P. 1995. p25rum1 orders S phase and mitosis by acting as an inhibitor of the
p34cdc2 mitotic kinase. Cell 83: 1001–1009.
Costanzo M, Nishikawa JL, Tang X, Millman JS, Schub O, Breitkreuz K, Dewar D, Rupes I, Andrews
B, Tyers M. 2004. CDK activity antagonizes Whi5, an inhibitor of G1/S transcription in yeast. Cell
117: 899–913.
Deb DK, Tanaka-Matakatsu M, Jones L, Richardson HE, Du W. 2008. Wingless signaling directly
regulates cyclin E expression in proliferating embryonic PNS precursor cells. Mech Dev 125: 857–
864.
de Bruin RA, McDonald WH, Kalashnikova TI, Yates J 3rd, Wittenberg C. 2004. Cln3 activates G1-
specific transcription via phosphorylation of the SBF bound repressor Whi5. Cell 117: 887–898.
Deng C, Zhang P, Harper JW, Elledge SJ, Leder P. 1995. Mice lacking p21CIP1/WAF1 undergo
normal development, but are defective in G1 checkpoint control. Cell 82: 675–684.
de Nooij JC, Letendre MA, Hariharan IK. 1996. A cyclin-dependent kinase inhibitor, Dacapo, is
necessary for timely exit from the cell cycle during Drosophila embryogenesis. Cell 87: 1237–
1247.
Diehl JA, Cheng M, Roussel M, Sherr CJ. 1998. Glycogen synthase kinase-3β regulates cyclin D1
proteolysis and subcellular localization. Genes Dev 12: 3499–3511.
Dulic V, Lees E, Reed SI. 1992. Association of human cyclin E with a periodic G1–S phase protein
kinase. Science 257: 1958–1961.
Duronio RJ, O’Farrell PH. 1995. Developmental control of the G1 to S transition in Drosophila: Cyclin
E is a limiting downstream target of E2F. Genes Dev 9: 1456–1468.
Dyson N. 1998. The regulation of E2F by pRB-family proteins. Genes Dev 12: 2245–2262.
El-Deiry WS, Tokino T, Velculescu VE, Levy DB, Parsons R, Lin DM, Mercer WE, Kinzler KWV,
Vogelstein B. 1993. WAF1, a potential mediator of p53 tumor suppression. Cell 75: 817–825.
Ewen ME, Sluss HK, Sherr CJ, Matsushime H, Kato J, Livingston DM. 1993. Functional interactions
of the retinoblastoma protein with mammalian D-type cyclins. Cell 73: 487–497.
Fay DS, Han M. 2000. Mutations in cye-1, a Caenorhabditis elegans cyclin E homolog, reveal
coordination between cell-cycle control and vulval development. Development 127: 4049–4060.
Feldman RMR, Correll CC, Kaplan KB, Deshaies RJ. 1997. A complex of Cdc4p, Skp1p, and
Cdc53p/Cullin catalyzes ubiquitination of the phosphorylated CDK inhibitor Sic1p. Cell 91: 221–
230.
Ferrell JE Jr, Pomerening JR, Kim SY, Trunnell NB, Xiong W, Huang CY, Machleder EM. 2009.
Simple, realistic models of complex biological processes: Positive feedback and bistability in a cell
fate switch and a cell cycle oscillator. FEBS Lett 583: 3999–4005.
Firth LC, Baker NE. 2005. Extracellular signals responsible for spatially regulated proliferation in the
differentiating Drosophila eye. Dev Cell 8: 541–551.
Franklin DS, Godfrey VL, Lee H, Kovalev GI, Schoonhoven R, Chen-Kiang S, Su L, Xiong Y. 1998.
CDK inhibitors p18INK4c and p27KIP1 mediate two separate pathways to collaboratively suppress
pituitary tumorigenesis. Genes Dev 12: 2899–2911.
Geng Y, Eaton EN, Picon M, Roberts JM, Lundberg AS, Gifford A, Sardet C, Weinberg RA. 1996.
Regulation of cyclin E transcription by E2Fs and retinoblastoma protein. Oncogene 12: 1173–1180.
Geng Y, Whoriskey W, Park MY, Bronson RT, Medema RH, Li T, Weinberg RA, Sicinski P. 1999.
Rescue of cyclin D1 deficiency by knockin cyclin E. Cell 97: 767–777.
Geng Y, Yu Q, Sicinska E, Das M, Schneider JE, Bhattacharya S, Rideout WM, Bronson RT, Gardner
H, Sicinski P. 2003. Cyclin E ablation in the mouse. Cell 114: 431–443.
Harper JW, Adami GR, Wei N, Keyomarsi K, Elledge SJ. 1993. The p21 Cdk-interacting protein Cip1
is a potent inhibitor of G1 cyclin-dependent kinases. Cell 75: 805–816.
* Harvey KF, Hariharan IK. 2012. The Hippo pathway. Cold Spring Harb Perspect Biol 4: a011288.
Havens CG, Walter JC. 2011. Mechanism of CRL4(Cdt2), a PCNA-dependent E3 ubiquitin ligase.
Genes Dev 25: 1568–1582.
* Hemmings BA, Restuccia DF. 2012. The PI3K-PKB/Akt pathway. Cold Spring Harb Perspect Biol
4: a011189.
Herrera RE, Sah VP, Williams BO, Makela TP, Weinberg RA, Jacks T. 1996. Altered cell cycle
kinetics, gene expression, and G1 restriction point regulation in Rb-deficient fibroblasts. Mol Cell
Biol 16: 2402–2407.
Hinds PW, Mittnacht S, Dulic V, Arnold A, Reed SI, Weinberg RA. 1992. Regulation of
retinoblastoma protein functions by ectopic expression of human cyclins. Cell 70: 993–1006.
Hong Y, Roy R, Ambros V. 1998. Developmental regulation of a cyclin-dependent kinase inhibitor
controls postembryonic cell cycle progression in Caenorhabditis elegans. Development 125: 3585–
3597.
Huang J, Wu S, Barrera J, Matthews K, Pan D. 2005. The Hippo signaling pathway coordinately
regulates cell proliferation and apoptosis by inactivating Yorkie, the Drosophila homolog of YAP.
Cell 122: 421–434.
Hwang HC, Clurman BE. 2005. Cyclin E in normal and neoplastic cell cycles. Oncogene 24: 2776–
2786.
Hwang CY, Lee C, Kwon KS. 2009. Extracellular signal-regulated kinase 2-dependent phosphorylation
induces cytoplasmic localization and degradation of p21Cip1. Mol Cell Biol 29: 3379–3389.
* Ingham PW. 2012. Hedgehog signaling. Cold Spring Harb Perspect Biol 4: a011221.
Inze D. 2005. Green light for the cell cycle. EMBO J 24: 657–662.
Jacobs JJ, Kieboom K, Marino S, DePinho RA, van Lohuizen M. 1999. The oncogene and Polycomb-
group gene bmi-1 regulates cell proliferation and senescence through the ink4a locus. Nature 397:
164–168.
Jones L, Richardson H, Saint R. 2000. Tissue-specific regulation of cyclin E transcription during
Drosophila melanogaster embryogenesis. Development 127: 4619–4630.
Kamura T, Hara T, Matsumoto M, Ishida N, Okumura F, Hatakeyama S, Yoshida M, Nakayama K,
Nakayama KI. 2004. Cytoplasmic ubiquitin ligase KPC regulates proteolysis of p27Kip1 at G1
phase. Nat Cell Biol 6: 1229–1235.
Kanie T, Onoyama I, Matsumoto A, Yamada M, Nakatsumi H, Tateishi Y, Yamamura S, Tsunematsu
R, Matsumoto M, Nakayama KI. 2012. Genetic reevaluation of the role of F-box proteins in cyclin
D1 degradation. Mol Cell Biol 32: 590–605.
Kato J-Y, Matsushime H, Hiebert SW, Ewen M, Sherr CJ. 1993. Direct binding of cyclin D to the
retinoblastoma gene product (pRb) and pRb phosphorylation by the cyclin D-dependent kinase
CDK4. Genes Dev 7: 331–342.
Kim M, Nakamoto T, Nishimori S, Tanaka K, Chiba T. 2008a. A new ubiquitin ligase involved in
p57KIP2 proteolysis regulates osteoblast cell differentiation. EMBO Rep 9: 878–884.
Kim Y, Starostina NG, Kipreos ET. 2008b. The CRL4Cdt2 ubiquitin ligase targets the degradation of
p21Cip1 to control replication licensing. Genes Dev 22: 2507–2519.
Kitagawa M, Higashi H, Jung HK, Suzuki-Takahashi I, Ikeda M, Tamai K, Kato J, Segawa K, Yoshida
E, Nishimura S, et al. 1996. The consensus motif for phosphorylation by cyclin D1–Cdk4 is
different from that for phosphorylation by cyclin A/E–Cdk2. EMBO J 15: 7060–7069.
Klein EA, Assoian RK. 2008. Transcriptional regulation of the cyclin D1 gene at a glance. J Cell Sci
121: 3853–3857.
Knoblich JA, Sauer K, Jones L, Richardson H, Saint R, Lehner CF. 1994. Cyclin E controls S phase
progression and its down-regulation during Drosophila embryogenesis is required for the arrest of
cell proliferation. Cell 77: 107–120.
Koff A, Giordano A, Desai D, Yamashita K, Harper JW, Elledge S, Nishimoto T, Morgan DO, Franza
BR, Roberts JM. 1992. Formation and activation of a cyclin E–cdk2 complex during the G1 phase
of the human cell cycle. Science 257: 1689–1694.
Kominami K-I, Toda T. 1997. Fission yeast WD-repeat protein Pop1 regulates genome ploidy through
ubiquitin–proteasome-mediated degradation of the CDK inhibitor Rum1 and the S-phase initiator
Cdc18. Genes Dev 11: 1548–1560.
* Kopan R. 2012. Notch signaling. Cold Spring Harb Perspect Biol 4: a011213.
Korzelius J, The I, Ruijtenberg S, Prinsen MB, Portegijs V, Middelkoop TC, Groot Koerkamp MJ,
Holstege FC, Boxem M, van den Heuvel S. 2011. Caenorhabditis elegans cyclin D/CDK4 and
cyclin E/CDK2 induce distinct cell cycle re-entry programs in differentiated muscle cells. PLoS
Genet 7: e1002362.
Kotake Y, Cao R, Viatour P, Sage J, Zhang Y, Xiong Y. 2007. pRB family proteins are required for
H3K27 trimethylation and Polycomb repression complexes binding to and silencing p16INK4a
tumor suppressor gene. Genes Dev 21: 49–54.
Kozar K, Ciemerych MA, Rebel VI, Shigematsu H, Zagozdzon A, Sicinska E, Geng Y, Yu Q,
Bhattacharya S, Bronson RT, et al. 2004. Mouse development and cell proliferation in the absence
of D-cyclins. Cell 118: 477–491.
Lane ME, Sauer K, Wallace K, Jan YN, Lehner CF, Vaessin H. 1996. Dacapo, a cyclin-dependent
kinase inhibitor, stops cell proliferation during Drosophila development. Cell 87: 1225–1235.
Lew D, Dulic V, Reed SI. 1991. Isolation of three novel human cyclins by rescue of G1 cyclin (Cln)
function in yeast. Cell 66: 1197–1206.
Liou YC, Ryo A, Huang HK, Lu PJ, Bronson R, Fujimori F, Uchida T, Hunter T, Lu KP. 2002. Loss of
Pin1 function in the mouse causes phenotypes resembling cyclin D1-null phenotypes. Proc Natl
Acad Sci 99: 1335–1340.
Liu TH, Li L, Vaessin H. 2002. Transcription of the Drosophila CKI gene dacapo is regulated by a
modular array of cis-regulatory sequences. Mech Dev 112: 25–36.
Malumbres M, Barbacid M. 2009. Cell cycle, CDKs and cancer: A changing paradigm. Nat Rev Cancer
9: 153–166.
Malumbres M, Sotillo R, Santamaria D, Galan J, Cerezo A, Ortega S, Dubus P, Barbacid M. 2004.
Mammalian cells cycle without the D-type cyclin-dependent kinases Cdk4 and Cdk6. Cell 118:
493–504.
Matsushime H, Roussel MF, Ashmum RA, Sherr CJ. 1991. Colony-stimulating factor 1 regulates a
novel gene (CYL1) during the G1 phase of the cell cycle. Cell 65: 701–713.
Merlo A, Herman JG, Mao L, Lee DJ, Gabrielson E, Burger PC, Baylin SB, Sidransky D. 1995. 5′ CpG
island methylation is associated with transcriptional silencing of the tumour suppressor
p16/CDKN2/MTS1 in human cancers. Nat Med 1: 686–692.
Meyer CA, Jacobs HW, Datar SA, Du W, Edgar BA, Lehner CF. 2000. Drosophila Cdk4 is required
for normal growth and is dispensable for cell cycle progression. EMBO J 19: 4533–4542.
Meyer CA, Kramer I, Dittrich R, Marzodko S, Emmerich J, Lehner CF. 2002. Drosophila p27Dacapo
expression during embryogenesis is controlled by a complex regulatory region independent of cell
cycle progression. Development 129: 319–328.
Minella AC, Loeb KR, Knecht A, Welcker M, Varnum-Finney BJ, Bernstein ID, Roberts JM, Clurman
BE. 2008. Cyclin E phosphorylation regulates cell proliferation in hematopoietic and epithelial
lineages in vivo. Genes Dev 22: 1677–1689.
* Morrison DK. 2012. MAP kinase pathways. Cold Spring Harb Perspect Biol 4: 011254.
Motokura T, Bloom T, Kim HG, Juppner H, Ruderman JV, Kronenberg HM, Arnold A. 1991. A novel
cyclin encoded by a bcl1-linked candidate oncogene. Nature 350: 512–515.
Nash P, Tang X, Orlicky S, Chen Q, Gertler FB, Mendenhall MD, Sicheri F, Pawson T, Tyers M. 2001.
Multisite phosphorylation of a CDK inhibitor sets a threshold for the onset of DNA replication.
Nature 414: 514–521.
Nicholson SC, Nicolay BN, Frolov MV, Moberg KH. 2011. Notch-dependent expression of the
archipelago ubiquitin ligase subunit in the Drosophila eye. Development 138: 251–260.
Nishitani H, Shiomi Y, Iida H, Michishita M, Takami T, Tsurimoto T. 2008. CDK inhibitor p21 is
degraded by a proliferating cell nuclear antigen-coupled Cul4–DDB1Cdt2 pathway during S phase
and after UV irradiation. J Biol Chem 283: 29045–29052.
* Nusse R. 2012. Wnt signaling. Cold Spring Harb Perspect Biol 4: a011163.
Ohtani K, Degregori J, Nevins JR. 1995. Regulation of the cyclin E gene by transcription factor E2F1.
Proc Natl Acad Sci 92: 12146–12150.
Ohtsubo M, Roberts JM. 1993. Cyclin-dependent regulation of G1 in mammalian fibroblasts. Science
259: 1908–1912.
Ortega S, Malumbres M, Barbacid M. 2002. Cyclin D-dependent kinases, INK4 inhibitors and cancer.
Biochim Biophys Acta 1602: 73–87.
Ortega S, Prieto I, Odajima J, Martin A, Dubus P, Sotillo R, Barbero JL, Malumbres M, Barbacid M.
2003. Cyclin-dependent kinase 2 is essential for meiosis but not for mitotic cell division in mice.
Nat Genet 35: 25–31.
Pagano M, Tam SW, Theodoras AM, Beer-Romero P, Del Sal G, Chau V, Yew PR, Draetta GF, Rolfe
M. 1995. Role of the ubiquitin–proteasome pathway in regulating abundance of the cyclin-
dependent kinase inhibitor p27. Science 269: 682–685.
Parisi T, Beck AR, Rougier N, McNeil T, Lucian L, Werb Z, Amati B. 2003. Cyclins E1 and E2 are
required for endoreplication in placental trophoblast giant cells. EMBO J 22: 4794–4803.
Pei XH, Bai F, Smith MD, Usary J, Fan C, Pai S-Y, Ho IC, Perou CM, Xiong Y. 2009. CDK inhibitor
p18INK4c is a downstream target of GATA3 and restrains mammary luminal progenitor cell
proliferation and tumorigenesis. Cancer Cell 15: 389–401.
Resnitzky D, Gossen M, Bujard H, Reed S. 1994. Acceleration of the G1/S phase transition by
expression of cyclins D1 and E with an inducible system. Mol Cell Biol 14: 1669–1679.
* Rhind N, Russel P. 2012. Signaling pathways that regulate cell division. Cold Spring Harb Perspect
Biol 4: a005942.
Roussel MF. 1999. The INK4 family of cell cycle inhibitors in cancer. Oncogene 18: 5311–5317.
Schneider BL, Yang QH, Futcher AB. 1996. Linkage of replication to start by the Cdk inhibitor Sic1.
Science 272: 560–562.
Schwob E, Bohm T, Mendenhall MD, Nasmyth K. 1994. The B-type cyclin kinase inhibitor p40SIC1
controls the G1 to S transition in S. cerevisiae. Cell 79: 233–244.
Serrano M, Hannon GJ, Beach D. 1993. A new regulatory motif in cell cycle control causing specific
inhibition of cyclin D/CDK4. Nature 366: 704–707.
Sherr CJ. 1996. Cancer cell cycle. Science 274: 1672–1677.
Sherr CJ, Roberts JM. 1995. Inhibitors of mammalian G1 cyclin-dependent kinases. Genes Dev 9:
1149–1163.
Sherr CJ, Roberts JM. 2004. Living with or without cyclins and cyclin-dependent kinases. Genes Dev
18: 2699–2711.
Skowyra D, Craig K, Tyers M, Elledge SJ, Harper JW. 1997. F-box proteins are receptors that recruit
phosphorylated substrates to the SCF ubiquitin-ligase complex. Cell 91: 209–219.
Smith AP, Henze M, Lee JA, Osborn KG, Keck JM, Tedesco D, Bortner DM, Rosenberg MP, Reed SI.
2006. Deregulated cyclin E promotes p53 loss of heterozygosity and tumorigenesis in the mouse
mammary gland. Oncogene 25: 7245–7259.
Starostina NG, Kipreos ET. 2012. Multiple degradation pathways regulate versatile CIP/KIP CDK
inhibitors. Trends Cell Biol 22: 33–41.
Terada Y, Tatsuka M, Jinno S, Okayama H. 1995. Requirement for tyrosine phosphorylation of Cdk4
in G1 arrest induced by ultraviolet irradiation. Nature 376: 358–362.
van den Heuvel S, Dyson NJ. 2008. Conserved functions of the pRB and E2F families. Nat Rev Mol
Cell Biol 9: 713–724.
van Lohuizen M, Verbeek S, Scheijen B, Wientjens E, van der Gulden H, Berns A. 1991. Identification
of cooperating oncogenes in Eμ-myc transgenic mice by provirus tagging. Cell 65: 737–752.
Wander SA, Zhao D, Slingerland JM. 2011. p27: A barometer of signaling deregulation and potential
predictor of response to targeted therapies. Clin Cancer Res 17: 12–18.
Wang TC, Cardiff RD, Zukerberg L, Lees E, Arnold A, Schmidt EV. 1994. Mammary hyperplasia and
carcinoma in MMTV–cyclin D1 transgenic mice. Nature 369: 669–671.
Weinberg RA. 1995. The retinoblastoma protein and cell cycle control. Cell 81: 323–330.
Welcker M, Clurman BE. 2008. FBW7 ubiquitin ligase: A tumour suppressor at the crossroads of cell
division, growth and differentiation. Nat Rev Cancer 8: 83–93.
Xiong W, Ferrell JE Jr. 2003. A positive-feedback-based bistable “memory module” that governs a cell
fate decision. Nature 426: 460–465.
Xiong Y, Connolly T, Futcher B, Beach D. 1991. Human D-type cyclin. Cell 65: 691–699.
Xiong Y, Zhang H, Beach D. 1992. D-type cyclins associate with multiple protein kinases and the
DNA replication and repair factor PCNA. Cell 71: 505–514.
Xiong Y, Hannon G, Zhang H, Casso D, Kobayashi R, Beach D. 1993a. p21 is a universal inhibitor of
the cyclin kinases. Nature 366: 701–704.
Xiong Y, Zhang H, Beach D. 1993b. Subunit rearrangement of cyclin-dependent kinases is associated
with cellular transformation. Genes Dev 7: 1572–1583.
Yu Q, Geng Y, Sicinski P. 2001. Specific protection against breast cancers by cyclin D1 ablation.
Nature 411: 1017–1021.
Yuan Y, Shen H, Franklin DS, Scadden DT, Cheng T. 2004. In vivo self-renewing divisions of
haematopoietic stem cells are increased in the absence of the early G1-phase inhibitor, p18INK4C.
Nat Cell Biol 6: 436–442.
Zhao H, Chen X, Gurian-West M, Roberts JM. 2012. Loss of cyclin-dependent kinase 2 (CDK2)
inhibitory phosphorylation in a CDK2AF knock-in mouse causes misregulation of DNA replication
and centrosome duplication. Mol Cell Biol 32: 1421–1432.
Zindy F, Quelle DE, Roussel MF, Sherr CJ. 1997. Expression of the p16INK4a tumor suppressor
versus other INK4 family members during mouse development and aging. Oncogene 15: 203–211.
5CDKs are a family of kinases that regulate the cell cycle and that require binding to noncatalytic
partner proteins termed cyclins for activity.
6Linked to p16, both structurally in the genome and through regulation by Polycomb group proteins, is
the product of the ARF tumor suppressor gene, which is transcribed from an alternative promoter and
translated in an alternative reading frame from p16. ARF does not share any amino acid sequence
similarity with INK4 proteins and instead acts as a p53 activator by binding to and inhibiting the
activity of MDM2, the principle E3 ubiquitin ligase for and negative regulator of p53. As a result, any
signal, such as oncogenic stimulation, that induces the expression of ARF will stabilize p53 and
activate p21, leading to G1 cell cycle arrest.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a008904
CHAPTER 6
SUMMARY
Outline
1 Introduction
2 CDK1 and entry into mitosis
3 Regulation of the G2/M transition
4 Checkpoint regulation of the G2/M transition
5 The DNA damage and replication checkpoints
6 The DNA damage checkpoint triggered by DSBs
7 The DNA replication checkpoint triggered by stalled replication forks
8 Chk1 and Chk2 target Cdc25
9 The MAPK-dependent stress checkpoint
10 Resolution of G2 checkpoint arrests
11 Regulation of mitosis
12 Dissecting the network
13 Anaphase entry
14 Checkpoint regulation of mitosis
15 Regulation of cytokinesis
16 Concluding remarks
References
1 INTRODUCTION
The cell cycle (see Fig. 1) consists of DNA synthesis (S) and mitosis (M)
phases separated by gap phases in the order G1–S–G2–M (Murray and Hunt
1993; Nurse 2000; Morgan 2006). Cell division involves two connected
processes triggered at the end of G2 phase: mitosis itself (segregation of the
chromosomes, which duplicate in S phase) and cytokinesis (division of the
cell, per se). Mitosis can be subdivided into six distinct phases (see Box 1):
(1) prophase, in which the spindle begins to assemble in the cytoplasm and
chromosomes begin to condense in the nucleus; (2) prometaphase, in which
the nuclear envelope breaks down and chromosomes attach to the spindle; (3)
metaphase, in which chromosomes align at the spindle midzone; (4) anaphase
A, in which chromosomes move to the centrosomes, which form the spindle
poles; (5) anaphase B, in which the spindle elongates; and (6) telophase, in
which the nuclear envelope reforms around the new daughter nuclei. Mitosis
in yeasts differs in that the nuclear envelope does not break down; instead the
spindle-pole body, the yeast equivalent of the centrosome, spans the nuclear
envelope, allowing the spindle to access both the nucleus and the cytoplasm.
Signals during telophase trigger cytokinesis, which separates the daughter
nuclei into two daughter cells.
Figure 1. The major events of the cell cycle. The major events of the cell cycle are regulated by
successive waves of kinase and ubiquitin ligase activity. G1-cyclin–CDK activity is required to initiate
the cell cycle and activate B-type-cyclin–CDK activity. Low levels of B-type-cyclin–CDK activity are
sufficient to trigger S phase, but tyrosine phosphorylation by Wee1 prevents full activation, preventing
premature mitosis. Full CDK activation triggers mitosis and activates APC, which triggers anaphase
and feeds back to inactivate CDK activity. Inactivation of CDK allows exit from mitosis and the
reestablishment of interphase chromosome and nuclear structure in G1 phase. See Box 1 for description
of the stages of mitosis.
BOX 1. THE MAIN STAGES OF MITOSIS
10 RESOLUTION OF G2 CHECKPOINT
ARRESTS
G2 checkpoint arrests are generally reversible. Once the problem is resolved,
the checkpoint is inactivated and the cell can proceed with mitosis. This
inactivation appears to be a passive process in which resolution of the
checkpoint-initiating event (e.g., the repair of the DNA damage) removes the
signal that activates the checkpoint kinases, allowing constitutive
phosphatases to remove the phosphates from their substrates and reset the
system. In situations in which the problem cannot be resolved, cells can adapt
to the checkpoint signal and enter mitosis in the presence of ongoing
checkpoint signaling. Such adaptation generally appears to be a failure of
checkpoint function, as opposed to a regulated attenuation of checkpoint
signaling (Harrison and Haber 2006). As cells continue to grow during a
checkpoint arrest, it is likely that the mitosis-promoting activities that link the
onset of mitosis to attainment of a sufficient cell size eventually overcome
mitosis-inhibiting activity of the checkpoint. Alternatively, in metazoans,
transcriptional circuits involving p53 and p21 can be activated to drive cells
into senescence or apoptosis in response to prolonged G2 arrest (Kastan and
Bartek 2004). A failure to activate these circuits can lead to premature
resumption of cell division in the presence of DNA damage (Bunz et al.
1998).
11 REGULATION OF MITOSIS
If the G2 checkpoints are not triggered, cells fully activate CDK1 and proceed
through the G2/M transition into mitosis. This decision to commit to cell
division is implemented by signaling pathways that regulate the various
processes of cell division, in particular mitosis and cytokinesis. They also
regulate organelles and other cytoplasmic components in ways that are less
well understood.
The transition from G2 phase to mitosis involves reorganization of the
nucleus, the condensation of the chromosomes, and the formation of the
mitotic spindle (see Box 1 and Fig. 3). These events are triggered by CDK1
and culminate with the mitotic chromosomes aligned on the metaphase plate
(Morgan 2006). How it coordinates these processes is a long-standing
question in the field.
13 ANAPHASE ENTRY
Once chromosomes are properly aligned on the metaphase plate, anaphase is
triggered via the activation of the APC (Fig. 3) (Morgan 2006). The APC is a
large multisubunit E3 ubiquitin ligase that regulates the stability of a range of
mitotic proteins by targeting them for ubiquitin-dependent proteolysis. It is
regulated by the binding of one of two substrate-selectivity subunits: Cdc20
at the metaphase/anaphase transition and Cdh1 during telophase and into G1
phase (Pesin and Orr-Weaver 2008). APC–Cdc20 promotes the degradation
of several key substrates that trigger the irreversible transition from
metaphase to anaphase and the subsequent exit from mitosis. One key
substrate is securin, a stoichiometric inhibitor of separase, the protease that
cleaves the cohesin complexes that hold sister chromatids together during
metaphase. Another key substrate is cyclin B, degradation of which
inactivates CDK1. In addition, APC activation leads to the indirect activation
of the Cdc14 phosphatase, which dephosphorylates many CDK1 targets,
further enforcing the inactivation of CDK1 (Clifford et al. 2008). PP1 and
PP2A also play a role (Wurzenberger and Gerlich 2011).
In yeasts, there appears to be an intrinsic delay between the activation of
CDK1 and the activation of APC, which usually gives the cells enough time
to set up the metaphase plate. APC is activated by CDK1 phosphorylation
(Kraft et al. 2003). The lag may give chromosomes adequate time to align. If
not, a checkpoint signal (described below) prevents APC activation until the
problem is resolved. In most metazoans, metaphase is rarely achieved in
time, presumably because the metazoan spindle is larger and more
complicated, and the checkpoint is activated in most cell cycles to delay
anaphase until the chromosomes are properly aligned. In effect, the regulation
of anaphase has gone from being a quality-control checkpoint in yeast to
being a central signaling pathway in metazoans that triggers anaphase upon
the successful completion of metaphase.
15 REGULATION OF CYTOKINESIS
The proper coordination of cytokinesis with mitosis is essential to ensure
faithful chromosome segregation and avoid aneuploidy or polyploidy. This
coordination requires the septation initiation network (SIN) in fission yeast
and the mitotic exit network (MEN) in budding yeast (McCollum and Gould
2001; Goyal et al. 2011; Meitinger et al. 2012). Cytokinesis is entrained to
mitosis; thus, the signaling pathways that regulate cytokinesis are not
involved so much in deciding when cytokinesis should occur as in
coordinating cytokinesis with mitosis and providing opportunities for
checkpoint regulation.
The MEN and SIN pathways are GTPase-triggered kinase cascades that
culminate in the activation of the Sid2 kinase in fission yeast and the Dbf2
kinase in budding yeast (Goyal et al. 2011; Meitinger et al. 2012).
Components of the MEN/SIN pathways are organized on the spindle-pole
body, the yeast equivalent of the centrosome, making it a nexus for signaling
pathways that control cell division. Sid2 is necessary and sufficient for
initiating cytokinesis in fission yeast, although its exact targets are not
known. Therefore, it is essential to restrain SIN signaling until after the
successful completion of anaphase. Initiation of SIN signaling by the Spg1
GTPase is inhibited by the GTPase-activating protein Cdc16. Full activation
of SIN signaling is antagonized by CDK1 activity (Guertin et al. 2000),
which prevents cytokinesis until after activation of the APC and ensures that
activation of the spindle assembly checkpoint will also delay cell division. In
budding yeast, Dbf2 regulates cytokinesis by promoting localization of the
chitin synthase Chs2 and the cytokinesis regulator Hof1 to the bud neck; this
activity is antagonized by CDK1 activity (Meitinger et al. 2012).
In return, the MEN/SIN pathways antagonize the activity of CDK1
(Goyal et al. 2011; Meitinger et al. 2012). This reciprocal regulation allows
cytokinesis errors, which prolong SIN signaling, to restrain CDK1 activity in
the subsequent cell cycle, thus arresting cells in G2 phase and preventing the
next mitosis until the previous cytokinesis is successfully completed. An
important component of the MEN/SIN pathways is the Cdc14 (Clp1 in
fission yeast) phosphatase (Clifford et al. 2008). Cdc14 directly
dephosphorylates CDK1 targets, facilitating mitotic exit and resetting the cell
to an interphase state at the beginning of G1 phase.
Many proteins in the MEN/SIN pathways are conserved in metazoans. In
particular, the LATS kinases, relatives of Sid2/Dbf2 that also function in the
Hippo pathway (p. 133 [Harvey and Hariharan 2012]), appear to have roles in
the regulation of cytokinesis (Yang et al. 2004). Although the details have yet
to be established, similar signaling pathways probably coordinate mitosis and
cytokinesis in metazoans.
In addition to coordinating the timing of cytokinesis, signaling during cell
division is required to determine the location of cytokinesis. Although the
details differ between organisms, the location and orientation of the
cytokinesis furrow is generally determined by the mitotic spindle, except in
fission yeast, in which the cleavage plane is determined directly by the
location of the nucleus (Almonacid and Paoletti 2010). How the signal is
transmitted from the spindle or the nucleus to the cortex to establish the site
of cytokinesis has yet to be established. Budding yeast is unusual in this
context because the cleavage plane (the bud neck) is established before
mitosis. Therefore, instead of using the spindle to orient cell division,
budding yeast uses cell division to orient the spindle. Specifically, the MEN
GTPase Tem1 is localized to the spindle-pole body, whereas Tem1’s
inhibitors, Bub2, Bfa1, and Kin4, are localized to the mother cell, and its
activator Lte1, a guanine nucleotide exchange factor (GEF) relative, is
localized to the daughter cell (Bardin et al. 2000). Thus, the MEN is only
activated once the spindle is oriented such that one end of the spindle is
through the bud neck and in the daughter cell. However, this strategy of
triggering cytokinesis as a spatial consequence of spindle elongation may be
general, as a similar mechanism functions in fission yeast (Garcia-Cortes and
McCollum 2009).
16 CONCLUDING REMARKS
The major cell-cycle transitions that constitute cell division—the G2/M
transition, the metaphase/anaphase transition, and cytokinesis—provide
important decision points that are regulated by a number of signaling
pathways. These pathways ensure that the critical events of cell division
occur in the proper order and provide the quality controls that prevent cells
from dividing with damaged DNA or misaligned chromosomes. As such,
they are instrumental in maintaining genomic integrity and, in metazoans,
preventing cancer.
These negative regulatory signaling pathways were the original
inspiration for the checkpoint paradigm of active negative regulation of cell-
cycle events (Hartwell and Weinert 1989). The initial model posited that
checkpoints delayed cell-cycle transitions to allow time for checkpoint-
independent processes to fix whatever problem had triggered the checkpoint.
Since then, these same pathways have been shown to regulate many other
aspects of cell metabolism, such as DNA repair, and the term “checkpoint” is
now used much more broadly than originally intended. Nonetheless, these
signaling pathways serve as prime examples of how cells reorganize their
metabolism and cell cycle to damage and other perturbations.
Although the well-studied signaling pathways that regulate cell division
inhibit transitions in response to signals of damage or other problems, there is
evidence for at least one positive signaling pathway, the one that regulates
cell size. One of the enduring mysteries of cell biology is how cells measure
size and how that information is used to regulate cell-cycle transitions such as
cell division. Notwithstanding the current lack of mechanistic insight, the
way size is measured and the pathways that transduce that signal are poised
to be areas of future progress in the field.
The signal transduction pathways that regulate cell division continue to be
the focus of significant experimental effort. That effort looks set to continue
as work continues on the discovery of new regulators of cell division, the
increasingly detailed mechanistic understanding of the major checkpoint
pathways, and the translation of our understanding of these pathways into
diagnostic and therapeutic advances in fields such as fertility, cancer, and
aging.
ACKNOWLEDGMENTS
We thank Dan McCollum for valuable insight. N.R. is supported by NIH
R01-GM069957 and an ACS Research Scholar Grant. P.R. is supported by
NIH R01-GM59447, CA77325, and CA117638.
REFERENCES
*Reference is in this book.
Almonacid M, Paoletti A. 2010. Mechanisms controlling division-plane positioning. Semin Cell Dev
Biol 21: 874–880.
Archambault V, Glover DM. 2009. Polo-like kinases: Conservation and divergence in their functions
and regulation. Nat Rev Mol Cell Biol 10: 265–275.
Bardin AJ, Visintin R, Amon A. 2000. A mechanism for coupling exit from mitosis to partitioning of
the nucleus. Cell 102: 21–31.
Barr FA, Sillje HH, Nigg EA. 2004. Polo-like kinases and the orchestration of cell division. Nat Rev
Mol Cell Biol 5: 429–440.
Bartek J, Lukas C, Lukas J. 2004. Checking on DNA damage in S phase. Nat Rev Mol Cell Biol 5:
792–804.
Beausoleil SA, Villen J, Gerber SA, Rush J, Gygi SP. 2006. A probability-based approach for high-
throughput protein phosphorylation analysis and site localization. Nat Biotechnol 24: 1285–1292.
Bloom J, Cross FR. 2007. Multiple levels of cyclin specificity in cell-cycle control. Nat Rev Mol Cell
Biol 8: 149–160.
Bonilla CY, Melo JA, Toczyski DP. 2008. Colocalization of sensors is sufficient to activate the DNA
damage checkpoint in the absence of damage. Mol Cell 30: 267–276.
Bonner WM, Redon CE, Dickey JS, Nakamura AJ, Sedelnikova OA, Solier S, Pommier Y. 2008.
γH2AX and cancer. Nat Rev Cancer 8: 957–967.
Branzei D, Foiani M. 2010. Maintaining genome stability at the replication fork. Nat Rev Mol Cell Biol
11: 208–219.
Bunz F, Dutriaux A, Lengauer C, Waldman T, Zhou S, Brown JP, Sedivy JM, Kinzler KW, Vogelstein
B. 1998. Requirement for p53 and p21 to sustain G2 arrest after DNA damage. Science 282: 1497–
1501.
Castilho PV, Williams BC, Mochida S, Zhao Y, Goldberg ML. 2009. The M phase kinase Greatwall
(Gwl) promotes inactivation of PP2A/B55δ, a phosphatase directed against CDK phosphosites. Mol
Biol Cell 20: 4777–4789.
Cimprich KA, Cortez D. 2008. ATR: An essential regulator of genome integrity. Nat Rev Mol Cell Biol
9: 616–627.
Clifford DM, Chen CT, Roberts RH, Feoktistova A, Wolfe BA, Chen JS, McCollum D, Gould KL.
2008. The role of Cdc14 phosphatases in the control of cell division. Biochem Soc Trans 36: 436–
438.
Coudreuse D, Nurse P. 2010. Driving the cell cycle with a minimal CDK control network. Nature 468:
1074–1079.
de Bruin RA, Kalashnikova TI, Aslanian A, Wohlschlegel J, Chahwan C, Yates JR, Russell P,
Wittenberg C. 2008. DNA replication checkpoint promotes G1-S transcription by inactivating the
MBF repressor Nrm1. Proc Natl Acad Sci 105: 11230–11235.
Domingo-Sananes MR, Kapuy O, Hunt T, Novak B. 2011. Switches and latches: A biochemical tug-of-
war between the kinases and phosphatases that control mitosis. Philos Trans R Soc Lond B Biol Sci
366: 3584–3594.
Du LL, Nakamura TM, Russell P. 2006. Histone modification-dependent and -independent pathways
for recruitment of checkpoint protein Crb2 to double-strand breaks. Genes Dev 20: 1583–1596.
* Duronio RJ, Xiong Y. 2012. Signaling pathways that control cell proliferation. Cold Spring Harb
Perspect Biol 4: a008904.
Dutta C, Patel PK, Rosebrock A, Oliva A, Leatherwood J, Rhind N. 2008. The DNA replication
checkpoint directly regulates MBF-dependent G1/S transcription. Mol Cell Biol 28: 5977–5985.
Fantes P, Nurse P. 1977. Control of cell size at division in fission yeast by a growth-modulated size
control over nuclear division. Exp Cell Res 107: 377–386.
Featherstone C, Russell P. 1991. Fission yeast p107wee1 mitotic inhibitor is a tyrosine/serine kinase.
Nature 349: 808–811.
Garcia-Cortes JC, McCollum D. 2009. Proper timing of cytokinesis is regulated by
Schizosaccharomyces pombe Etd1. J Cell Biol 186: 739–753.
Gautier J, Solomon MJ, Booher RN, Bazan JF, Kirschner MW. 1991. cdc25 is a specific tyrosine
phosphatase that directly activates p34cdc2. Cell 67: 197–211.
Gharbi-Ayachi A, Labbe JC, Burgess A, Vigneron S, Strub JM, Brioudes E, Van-Dorsselaer A, Castro
A, Lorca T. 2010. The substrate of Greatwall kinase, Arpp19, controls mitosis by inhibiting protein
phosphatase 2A. Science 330: 1673–1677.
Gould KL, Nurse P. 1989. Tyrosine phosphorylation of the fission yeast cdc2+ protein kinase regulates
entry into mitosis. Nature 342: 39–45.
Goyal A, Takaine M, Simanis V, Nakano K. 2011. Dividing the spoils of growth and the cell cycle:
The fission yeast as a model for the study of cytokinesis. Cytoskeleton (Hoboken) 68: 69–88.
Guertin DA, Chang L, Irshad F, Gould KL, McCollum D. 2000. The role of the Sid1p kinase and
Cdc14p in regulating the onset of cytokinesis in fission yeast. EMBO J 19: 1803–1815.
Haase SB, Reed SI. 1999. Evidence that a free-running oscillator drives G1 events in the budding yeast
cell cycle. Nature 401: 394–397.
Hachet O, Berthelot-Grosjean M, Kokkoris K, Vincenzetti V, Moosbrugger J, Martin SG. 2011. A
phosphorylation cycle shapes gradients of the DYRK family kinase Pom1 at the plasma membrane.
Cell 145: 1116–1128.
Harper JW, Elledge SJ. 2007. The DNA damage response: Ten years after. Mol Cell 28: 739–745.
Harrison JC, Haber JE. 2006. Surviving the breakup: The DNA damage checkpoint. Annu Rev Genet
40: 209–235.
Hartwell LH, Weinert TA. 1989. Checkpoints: Controls that ensure the order of cell cycle events.
Science 246: 629–634.
* Harvey KF, Hariharan IK. 2012. Hippo signaling. Cold Spring Harb Perspect Biol 4: a011288.
* Hotamisligil GS, Davis RJ. 2014. Cell signaling and stress responses. Cold Spring Harb Perspect
Biol doi: 10.1101/cshperspect.a006072.
Howell AS, Lew DJ. 2012. Morphogenesis and the cell cycle. Genetics 190: 51–77.
Jackson SP, Bartek J. 2009. The DNA-damage response in human biology and disease. Nature 461:
1071–1078.
Jorgensen P, Tyers M. 2004. How cells coordinate growth and division. Curr Biol 14: R1014–R1027.
Kalaszczynska I, Geng Y, Iino T, Mizuno S, Choi Y, Kondratiuk I, Silver DP, Wolgemuth DJ, Akashi
K, Sicinski P. 2009. Cyclin A is redundant in fibroblasts but essential in hematopoietic and
embryonic stem cells. Cell 138: 352–365.
Karlsson-Rosenthal C, Millar JB. 2006. Cdc25: Mechanisms of checkpoint inhibition and recovery.
Trends Cell Biol 16: 285–292.
Kastan MB, Bartek J. 2004. Cell-cycle checkpoints and cancer. Nature 432: 316–323.
Kellogg DR. 2003. Wee1-dependent mechanisms required for coordination of cell growth and cell
division. J Cell Sci 116: 4883–4890.
Khodjakov A, Pines J. 2010. Centromere tension: A divisive issue. Nat Cell Biol 12: 919–923.
Kraft C, Herzog F, Gieffers C, Mechtler K, Hagting A, Pines J, Peters JM. 2003. Mitotic regulation of
the human anaphase-promoting complex by phosphorylation. EMBO J 22: 6598–6609.
Kumagai A, Dunphy WG. 1991. The cdc25 protein controls tyrosine dephosphorylation of the cdc2
protein in a cell-free system. Cell 64: 903–914.
Kumagai A, Dunphy WG. 1996. Purification and molecular cloning of Plx1, a Cdc25-regulatory kinase
from Xenopus egg extracts. Science 273: 1377–1380.
Kumagai A, Dunphy WG. 2000. Claspin, a novel protein required for the activation of Chk1 during a
DNA replication checkpoint response in Xenopus egg extracts. Mol Cell 6: 839–849.
* Laplante M, Sabatini DM. 2012. mTOR signaling. Cold Spring Harb Perspect Biol 4: a011593.
Li X, Nicklas RB. 1995. Mitotic forces control a cell-cycle checkpoint. Nature 373: 630–632.
Lundgren K, Walworth N, Booher R, Dembski M, Kirschner M, Beach D. 1991. Mik1 and Wee1
cooperate in the inhibitory tyrosine phosphorylation of Cdc2. Cell 64: 1111–1122.
Ma HT, Poon RY. 2011. How protein kinases co-ordinate mitosis in animal cells. Biochem J 435: 17–
31.
Martin SG, Berthelot-Grosjean M. 2009. Polar gradients of the DYRK-family kinase Pom1 couple cell
length with the cell cycle. Nature 459: 852–856.
Matsuoka S, Ballif BA, Smogorzewska A, McDonald ER III, Hurov KE, Luo J, Bakalarski CE, Zhao
Z, Solimini N, Lerenthal Y, et al. 2007. ATM and ATR substrate analysis reveals extensive protein
networks responsive to DNA damage. Science 316: 1160–1166.
McCollum D, Gould KL. 2001. Timing is everything: Regulation of mitotic exit and cytokinesis by the
MEN and SIN. Trends Cell Biol 11: 89–95.
McFarlane RJ, Mian S, Dalgaard JZ. 2010. The many facets of the Tim-Tipin protein families’ roles in
chromosome biology. Cell Cycle 9: 700–705.
Meitinger F, Palani S, Pereira G. 2012. The power of MEN in cytokinesis. Cell Cycle 11: 219–228.
Melander F, Bekker-Jensen S, Falck J, Bartek J, Mailand N, Lukas J. 2008. Phosphorylation of SDT
repeats in the MDC1 N terminus triggers retention of NBS1 at the DNA damage-modified
chromatin. J Cell Biol 181: 213–226.
Mimitou EP, Symington LS. 2011. DNA end resection—unraveling the tail. DNA Repair (Amst) 10:
344–348.
Morgan DO. 1995. Principles of CDK regulation. Nature 374: 131–134.
Morgan DO. 2006. The cell cycle: Principles of control. Oxford University Press, Sunderland, MA.
* Morrison DK. 2012. MAP kinase pathways. Cold Spring Harb Perspect Biol 4: a011254.
Moseley JB, Mayeux A, Paoletti A, Nurse P. 2009. A spatial gradient coordinates cell size and mitotic
entry in fission yeast. Nature 459: 857–860.
Mueller PR, Coleman TR, Kumagai A, Dunphy WG. 1995. Myt1: A membrane-associated inhibitory
kinase that phosphorylates Cdc2 on both threonine-14 and tyrosine-15. Science 270: 86–90.
Murray AW, Hunt T. 1993. The cell cycle: An introduction. W.H. Freeman, New York.
Musacchio A, Salmon ED. 2007. The spindle-assembly checkpoint in space and time. Nat Rev Mol Cell
Biol 8: 379–393.
Navadgi-Patil VM, Burgers PM. 2009. A tale of two tails: Activation of DNA damage checkpoint
kinase Mec1/ATR by the 9-1-1 clamp and by Dpb11/TopBP1. DNA Repair (Amst) 8: 996–1003.
Nurse P. 1975. Genetic control of cell size at cell division in yeast. Nature 256: 547–551.
Nurse P. 1990. Universal control mechanism regulating onset of M-phase. Nature 344: 503–508.
Nurse P. 2000. A long twentieth century of the cell cycle and beyond. Cell 100: 71–78.
Parker LL, Piwnica-Worms H. 1992. Inactivation of the p34cdc2–cyclin B complex by the human
WEE1 tyrosine kinase. Science 257: 1955–1957.
Pesin JA, Orr-Weaver TL. 2008. Regulation of APC/C activators in mitosis and meiosis. Annu Rev Cell
Dev Biol 24: 475–499.
Petersen J, Nurse P. 2007. TOR signalling regulates mitotic commitment through the stress MAP
kinase pathway and the Polo and Cdc2 kinases. Nat Cell Biol 9: 1263–1272.
Polymenis M, Schmidt EV. 1997. Coupling of cell division to cell growth by translational control of
the G1 cyclin CLN3 in yeast. Genes Dev 11: 2522–2531.
Pomerening JR, Sontag ED, Ferrell JE Jr. 2003. Building a cell cycle oscillator: Hysteresis and
bistability in the activation of Cdc2. Nat Cell Biol 5: 346–351.
Porter LA, Donoghue DJ. 2003. Cyclin B1 and CDK1: Nuclear localization and upstream regulators.
Prog Cell Cycle Res 5: 335–347.
Rhind N, Russell P. 2000. Chk1 and Cds1: Linchpins of the DNA damage and replication checkpoint
pathways. J Cell Sci 113: 3889–3896.
Rieder CL. 2011. Mitosis in vertebrates: The G2/M and M/A transitions and their associated
checkpoints. Chromosome Res 19: 291–306.
Rieder CL, Cole RW, Khodjakov A, Sluder G. 1995. The checkpoint delaying anaphase in response to
chromosome monoorientation is mediated by an inhibitory signal produced by unattached
kinetochores. J Cell Biol 130: 941–948.
Russell P, Nurse P. 1986. cdc25+ functions as an inducer in the mitotic control of fission yeast. Cell 45:
145–153.
Russell P, Nurse P. 1987. Negative regulation of mitosis by wee1+, a gene encoding a protein kinase
homolog. Cell 49: 559–567.
Sanchez Y, Bachant J, Wang H, Hu F, Liu D, Tetzlaff M, Elledge SJ. 1999. Control of the DNA
damage checkpoint by Chk1 and Rad53 protein kinases through distinct mechanisms. Science 286:
1166–1171.
Shiloh Y. 2003. ATM and related protein kinases: Safeguarding genome integrity. Nat Rev Cancer 3:
155–168.
Shiozaki K, Russell P. 1995. Cell-cycle control linked to extracellular environment by MAP kinase
pathway in fission yeast. Nature 378: 739–743.
Spycher C, Miller ES, Townsend K, Pavic L, Morrice NA, Janscak P, Stewart GS, Stucki M. 2008.
Constitutive phosphorylation of MDC1 physically links the MRE11–RAD50–NBS1 complex to
damaged chromatin. J Cell Biol 181: 227–240.
Stern B, Nurse P. 1996. A quantitative model for the cdc2 control of S phase and mitosis in fission
yeast. Trends Genet 12: 345–350.
Stewart GS, Wang B, Bignell CR, Taylor AM, Elledge SJ. 2003. MDC1 is a mediator of the
mammalian DNA damage checkpoint. Nature 421: 961–966.
Stracker TH, Petrini JH. 2011. The MRE11 complex: Starting from the ends. Nat Rev Mol Cell Biol 12:
90–103.
Strausfeld U, Labbe JC, Fesquet D, Cavadore JC, Picard A, Sadhu K, Russell P, Doree M. 1991.
Dephosphorylation and activation of a p34cdc2/cyclin B complex in vitro by human CDC25
protein. Nature 351: 242–245.
Stucki M, Clapperton JA, Mohammad D, Yaffe MB, Smerdon SJ, Jackson SP. 2005. MDC1 directly
binds phosphorylated histone H2AX to regulate cellular responses to DNA double-strand breaks.
Cell 123: 1213–1226.
Tanaka TU, Rachidi N, Janke C, Pereira G, Galova M, Schiebel E, Stark MJ, Nasmyth K. 2002.
Evidence that the Ipl1-Sli15 (Aurora kinase-INCENP) complex promotes chromosome bi-
orientation by altering kinetochore-spindle pole connections. Cell 108: 317–329.
Tyson JJ. 1983. Unstable activator models for size control of the cell cycle. J Theor Biol 104: 617–631.
Tzur A, Kafri R, LeBleu VS, Lahav G, Kirschner MW. 2009. Cell growth and size homeostasis in
proliferating animal cells. Science 325: 167–171.
Uhlmann F, Bouchoux C, Lopez-Aviles S. 2011. A quantitative model for cyclin-dependent kinase
control of the cell cycle: Revisited. Philos Trans R Soc Lond B Biol Sci 366: 3572–3583.
* Ward PS, Thompson CB. 2012. Signaling in control of cell growth and metabolism. Cold Spring
Harb Perspect Biol 4: a006783.
Welburn JP, Tucker JA, Johnson T, Lindert L, Morgan M, Willis A, Noble ME, Endicott JA. 2007.
How tyrosine 15 phosphorylation inhibits the activity of cyclin-dependent kinase 2-cyclin A. J Biol
Chem 282: 3173–3181.
Wurzenberger C, Gerlich DW. 2011. Phosphatases: Providing safe passage through mitotic exit. Nat
Rev Mol Cell Biol 12: 469–482.
Yang X, Yu K, Hao Y, Li DM, Stewart R, Insogna KL, Xu T. 2004. LATS1 tumour suppressor affects
cytokinesis by inhibiting LIMK1. Nat Cell Biol 6: 609–617.
You Z, Shi LZ, Zhu Q, Wu P, Zhang YW, Basilio A, Tonnu N, Verma IM, Berns MW, Hunter T. 2009.
CtIP links DNA double-strand break sensing to resection. Mol Cell 36: 954–969.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a005942
CHAPTER 7
SUMMARY
Outline
1 Introduction
2 PI3K/AKT signaling controls glucose metabolism and the
incorporation of carbon into macromolecules
3 HIF1 signaling provides additional regulation of glucose metabolism
in response to both oxygen availability and nutrient status
4 Tyrosine kinase signaling and pyruvate kinase: Regulation of a
metabolic switch in proliferating cells
5 Amino acid metabolism is also regulated by cellular signaling
cascades
6 Metabolically sensitive protein modifications link growth factor
signaling to cellular responses
7 Perturbations of cellular metabolism in disease
8 Concluding remarks
References
1 INTRODUCTION
Unicellular organisms have evolved to grow and divide rapidly when
nutrients are abundant, and they take up nutrients in a cell-autonomous
manner. The macromolecular precursors and free energy derived from
metabolism of these nutrients are used to synthesize the new biomass
required for cell growth and division. When the nutrient supply dwindles,
anabolic metabolism in these organisms decreases. The cells then shift to
catabolic pathways that maximize the efficiency of energy production to
survive periods of nutrient limitation (Vander Heiden et al. 2009).
In multicellular organisms, cells are generally surrounded by sufficient
nutrients to engage in continuous cell growth and proliferation. However,
organismal integrity requires that proliferation not be a cell-autonomous
process dictated by available nutrients. Mammalian cells require receptor-
mediated signal transduction initiated by extracellular growth factors to leave
the quiescent state and enter the cell cycle. The onset of cell growth and
division introduces a metabolic requirement for sufficient carbon, nitrogen,
and free energy to support synthesis of the new proteins, lipids, and nucleic
acids needed by a proliferating cell. Recent studies have shown that this
additional uptake of nutrients is regulated by signal transduction pathways
(Fig. 1). This growth-factor-directed uptake of nutrients is critical to
supporting a rate of macromolecular synthesis sufficient for growth
(DeBerardinis et al. 2008; Vander Heiden et al. 2009).
Figure 1. Growth-factor-initiated signaling reprograms metabolism in proliferating cells. (A) In
multicellular organisms, cells that are not instructed to proliferate by extracellular growth factors are
generally quiescent. In these cells, glucose carbon is predominantly metabolized to carbon dioxide in
the mitochondrial Krebs cycle when oxygen is available. This mitochondrial oxidation maximizes free-
energy generation in the form of ATP. (B) When cells are instructed to proliferate by growth factor
signaling, they increase their nutrient uptake, particularly that of glucose and glutamine. Much of this
increased nutrient uptake is used to fulfill the lipid, protein, and nucleotide synthesis (biomass) required
for cell growth, and the excess carbon is secreted as lactate. Proliferating cells also may adopt strategies
to increase their ATP consumption to maintain glycolytic flux. Metabolic pathways are indicated by
green arrows.
The full range of effects that PKM2 may have on cell growth and
metabolism, however, remains incompletely characterized. Recently, an
alternative glycolytic pathway has been proposed to be present in PKM2-
expressing cells (Vander Heiden et al. 2010). When inactive PKM2 cannot
effectively convert PEP to pyruvate, PEP can donate its high-energy
phosphate to H11 of the upstream glycolytic enzyme phosphoglycerate
mutase. This pathway may allow the generation of pyruvate from PEP in a
step that is independent of ATP generation. Although further work is needed
to characterize and purify an enzyme that can catalyze this proposed activity,
the pathway may serve as an additional mechanism, besides Akt-mediated
up-regulation of ENTPD5, by which proliferating cells can decrease their
ATP:ADP ratio. The potential importance of PKM2 as a binding partner for a
variety of signaling kinases and transcription factors has also received
attention. For example, PKM2 translocates to the nucleus and potentiates the
transcriptional activity of the Oct4 transcription factor involved in
maintaining pluripotency (Lee et al. 2008). Many additional PKM2-
interacting proteins have been reported, most recently HIF1 and β-catenin,
but the significance of each of these interactions for facilitating cell growth
and proliferation requires further study.
The regulation of PKM2 activity by metabolites is also continuing to be
investigated. For example, ROS have recently been proposed to cause
oxidation, dissociation, and inactivation of the PKM2 tetramer (Anastasiou et
al. 2011). Conversely, the non-essential amino acid serine allosterically
activates PKM2 (Eigenbrodt et al. 1983). Ongoing work is addressing
whether inactivation of PKM2 and the resultant accumulation of glycolytic
intermediates can lead to enhanced flux through the serine synthetic pathway
and thereby promote a feedback loop through serine synthesis that can
modulate the flux of glycolytic intermediates to support cellular biosynthetic
requirements.
8 CONCLUDING REMARKS
In multicellular organisms, cell growth and proliferation are normally not cell
autonomous. Receptor-mediated signal transduction, initiated by extracellular
growth factors, promotes entry into the cell cycle and reprograms cellular
metabolism to fulfill the biosynthetic needs of cell growth and division (Fig.
6). However, despite having become highly dependent on instruction from
extracellular growth factors, mammalian cells have retained the ability to
sense their internal metabolic reserves and adjust their growth and
biosynthetic activities accordingly. Much of this feedback control occurs at
the level of posttranslational modifications of signal transduction proteins by
key cellular metabolites. Moreover, intracellular metabolites can also regulate
chromatin accessibility to control gene expression.
Figure 6. Growth factor signaling reprograms cellular metabolism to promote biosynthesis but is
sensitive to feedback control by metabolically sensitive protein modifications. Signaling downstream
from PI3K/Akt enhances glucose uptake, glycolysis, and the flux of glucose carbon into cytosolic
acetyl-CoA and lipids. HIF1 signaling further promotes glycolysis while also enhancing the flux of
pyruvate into lactate under conditions of O2 limitation or nutrient excess. The Myc transcription factor,
activated downstream from Ras, enhances glutamine uptake and metabolism as well as nucleotide
biosynthesis. mTORC1 signaling responds to both upstream PI3K/Akt signaling and the levels of
essential amino acids to promote protein synthesis. PKM2 is a pyruvate kinase isoform specific to
proliferating cells that is uniquely sensitive to inhibition by tyrosine kinase signaling downstream from
growth factors. Metabolically sensitive protein modifications, such as receptor glycosylation and
nuclear histone acetylation, provide a way for the cell to exert feedback control on the output of growth
factor signaling.
The evolution of the ability to regulate chromatin accessibility by specific
metabolites may have preceded the ability of growth factor signaling to
reprogram metabolism in multicellular organisms. Nonetheless, many aspects
of how variations in cellular metabolism influence chromatin accessibility
remain to be fully characterized. In addition to intracellular glucose
metabolism directly altering the acetylation state of histones, the methylation
state of histone lysine and DNA cytosine may also be metabolically
responsive. These methylation states were once thought to be irreversible, but
recent work has described demethylating enzymes for both histones and
DNA that are sustained by α-ketoglutarate production downstream from
glutamine metabolism. As discussed above, these α-ketoglutarate-dependent
enzymes can be inhibited by the 2HG produced by IDH1/IDH2 mutations.
Another area of continuing investigation is how minimizing ATP
production and enhancing ATP consumption can facilitate the rapid nutrient
metabolism of proliferating cells (Israelsen and Vander Heiden 2010).
Minimizing cellular ATP accumulation runs counter to the metabolic strategy
of quiescent cells, which completely oxidize the majority of glucose carbon
in the mitochondria to maximize ATP production. But proliferating cells with
activated growth factor signaling pathways usually take up nutrients far in
excess of the level required to maintain ATP levels and avoid AMPK
activation. Moreover, glycolysis in proliferating cells is limited by the rate of
ATP consumption, not ATP production (Scholnick et al. 1973). Increasing
cellular ATP consumption and the ADP:ATP ratio may be critical for
relieving the inhibition of glycolytic enzymes that can occur when ATP
levels are high, inhibition that would otherwise prevent high glycolytic flux
and enhanced macromolecular biosynthesis from glycolytic intermediates.
REFERENCES
*Reference is in this book.
Amary MF, Bacsi K, Maggiani F, Damato S, Halai D, Berisha F, Pollock R, O’Donnell P, Grigoriadis
A, Diss T, et al. 2011. IDH1 and IDH2 mutations are frequent events in central chondrosarcoma
and central and periosteal chondromas but not in other mesenchymal tumours. J Pathol 224: 334–
343.
Anastasiou D, Poulogiannis G, Asara JM, Boxer MB, Jiang JK, Shen M, Bellinger G, Sasaki AT,
Locasale JW, Auld DS, et al. 2011. Inhibition of pyruvate kinase M2 by reactive oxygen species
contributes to antioxidant responses. Science 334: 1278–1283.
Bauer DE, Harris MH, Plas DR, Lum JJ, Hammerman PS, Rathmell JC, Riley JL, Thompson CB.
2004. Cytokine stimulation of aerobic glycolysis in hematopoietic cells exceeds proliferative
demand. FASEB J 18: 1303–1305.
Bauer DE, Hatzivassiliou G, Zhao F, Andreadis C, Thompson CB. 2005. ATP citrate lyase is an
important component of cell growth and transformation. Oncogene 24: 6314–6322.
Berwick DC, Hers I, Heesom KJ, Moule SK, Tavare JM. 2002. The identification of ATP-citrate lyase
as a protein kinase B (Akt) substrate in primary adipocytes. J Biol Chem 277: 33895–33900.
Bobrovnikova-Marjon E, Hatzivassiliou G, Grigoriadou C, Romero M, Cavener DR, Thompson CB,
Diehl JA. 2008. PERK-dependent regulation of lipogenesis during mouse mammary gland
development and adipocyte differentiation. Proc Natl Acad Sci 105: 16314–16319.
Cai L, Tu BP. 2011. On acetyl-CoA as a gauge of cellular metabolic state. Cold Spring Harb Symp
Quant Biol doi: 10.1101/sqb.2011.76.010769.
Cai L, Sutter BM, Li B, Tu BP. 2011. Acetyl-CoA induces cell growth and proliferation by promoting
the acetylation of histones at growth genes. Mol Cell 42: 426–437.
Chowdhury R, Yeoh KK, Tian YM, Hillringhaus L, Bagg EA, Rose NR, Leung IK, Li XS, Woon EC,
Yang M, et al. 2011. The oncometabolite 2-hydroxyglutarate inhibits histone lysine demethylases.
EMBO Rep 12: 463–469.
Christofk HR, Vander Heiden MG, Harris MH, Ramanathan A, Gerszten RE, Wei R, Fleming MD,
Schreiber SL, Cantley LC. 2008a. The M2 splice isoform of pyruvate kinase is important for cancer
metabolism and tumour growth. Nature 452: 230–233.
Christofk HR, Vander Heiden MG, Wu N, Asara JM, Cantley LC. 2008b. Pyruvate kinase M2 is a
phosphotyrosine-binding protein. Nature 452: 181–186.
Clower CV, Chatterjee D, Wang Z, Cantley LC, Vander Heiden MG, Krainer AR. 2010. The
alternative splicing repressors hnRNP A1/A2 and PTB influence pyruvate kinase isoform
expression and cell metabolism. Proc Natl Acad Sci 107: 1894–1899.
Corradetti MN, Inoki K, Bardeesy N, DePinho RA, Guan KL. 2004. Regulation of the TSC pathway by
LKB1: Evidence of a molecular link between tuberous sclerosis complex and Peutz-Jeghers
syndrome. Genes Dev 18: 1533–1538.
Dang L, White DW, Gross S, Bennett BD, Bittinger MA, Driggers EM, Fantin VR, Jang HG, Jin S,
Keenan MC, et al. 2009. Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature
462: 739–744.
David CJ, Chen M, Assanah M, Canoll P, Manley JL. 2010. HnRNP proteins controlled by c-Myc
deregulate pyruvate kinase mRNA splicing in cancer. Nature 463: 364–368.
DeBerardinis RJ, Mancuso A, Daikhin E, Nissim I, Yudkoff M, Wehrli S, Thompson CB. 2007.
Beyond aerobic glycolysis: Transformed cells can engage in glutamine metabolism that exceeds the
requirement for protein and nucleotide synthesis. Proc Natl Acad Sci 104: 19345–19350.
DeBerardinis RJ, Lum JJ, Hatzivassiliou G, Thompson CB. 2008. The biology of cancer: Metabolic
reprogramming fuels cell growth and proliferation. Cell Metab 7: 11–20.
Deprez J, Vertommen D, Alessi DR, Hue L, Rider MH. 1997. Phosphorylation and activation of heart
6-phosphofructo-2-kinase by protein kinase B and other protein kinases of the insulin signaling
cascades. J Biol Chem 272: 17269–17275.
Eigenbrodt E, Leib S, Kramer W, Friis RR, Schoner W. 1983. Structural and kinetic differences
between the M2 type pyruvate kinases from lung and various tumors. Biomed Biochim Acta 42:
S278–S282.
Elstrom RL, Bauer DE, Buzzai M, Karnauskas R, Harris MH, Plas DR, Zhuang H, Cinalli RM, Alavi
A, Rudin CM, et al. 2004. Akt stimulates aerobic glycolysis in cancer cells. Cancer Res 64: 3892–
3899.
Fang M, Shen Z, Huang S, Zhao L, Chen S, Mak TW, Wang X. 2010. The ER UDPase ENTPD5
promotes protein N-glycosylation, the Warburg effect, and proliferation in the PTEN pathway. Cell
143: 711–724.
Figueroa ME, Abdel-Wahab O, Lu C, Ward PS, Patel J, Shih A, Li Y, Bhagwat N, Vasanthakumar A,
Fernandez HF, et al. 2010. Leukemic IDH1 and IDH2 mutations result in a hypermethylation
phenotype, disrupt TET2 function, and impair hematopoietic differentiation. Cancer Cell 18: 553–
567.
Gao P, Tchernyshyov I, Chang TC, Lee YS, Kita K, Ochi T, Zeller KI, De Marzo AM, Van Eyk JE,
Mendell JT, et al. 2009. c-Myc suppression of miR-23a/b enhances mitochondrial glutaminase
expression and glutamine metabolism. Nature 458: 762–765.
Gottlob K, Majewski N, Kennedy S, Kandel E, Robey RB, Hay N. 2001. Inhibition of early apoptotic
events by Akt/PKB is dependent on the first committed step of glycolysis and mitochondrial
hexokinase. Genes Dev 15: 1406–1418.
Gwinn DM, Shackelford DB, Egan DF, Mihaylova MM, Mery A, Vasquez DS, Turk BE, Shaw RJ.
2008. AMPK phosphorylation of raptor mediates a metabolic checkpoint. Mol Cell 30: 214–226.
Haigis MC, Sinclair DA. 2010. Mammalian sirtuins: Biological insights and disease relevance. Annu
Rev Pathol 5: 253–295.
Hara K, Yonezawa K, Kozlowski MT, Sugimoto T, Andrabi K, Weng QP, Kasuga M, Nishimoto I,
Avruch J. 1997. Regulation of eIF-4E BP1 phosphorylation by mTOR. J Biol Chem 272: 26457–
26463.
* Hardie DG. 2012. Organismal carbohydrate and lipid homeostasis. Cold Spring Harb Perspect Biol
4: a006031.
* Hemmings BA, Restuccia DF. 2012. The PI3K-PKB/Akt pathway. Cold Spring Harb Perspect Biol
4: a011189.
Hitosugi T, Kang S, Vander Heiden MG, Chung TW, Elf S, Lythgoe K, Dong S, Lonial S, Wang X,
Chen GZ, et al. 2009. Tyrosine phosphorylation inhibits PKM2 to promote the Warburg effect and
tumor growth. Sci Signal 2: ra73.
Hudson CC, Liu M, Chiang GG, Otterness DM, Loomis DC, Kaper F, Giaccia AJ, Abraham RT. 2002.
Regulation of hypoxia-inducible factor 1α expression and function by the mammalian target of
rapamycin. Mol Cell Biol 22: 7004–7014.
* Ingham P. 2012. Hedgehog signaling. Cold Spring Harb Perspect Biol 4: a011221.
Isaacs JS, Jung YJ, Mole DR, Lee S, Torres-Cabala C, Chung YL, Merino M, Trepel J, Zbar B, Toro J,
et al. 2005. HIF overexpression correlates with biallelic loss of fumarate hydratase in renal cancer:
Novel role of fumarate in regulation of HIF stability. Cancer Cell 8: 143–153.
Israelsen WJ, Vander Heiden MG. 2010. ATP consumption promotes cancer metabolism. Cell 143:
669–671.
Kim JW, Tchernyshyov I, Semenza GL, Dang CV. 2006. HIF-1-mediated expression of pyruvate
dehydrogenase kinase: A metabolic switch required for cellular adaptation to hypoxia. Cell Metab
3: 177–185.
Kohn AD, Summers SA, Birnbaum MJ, Roth RA. 1996. Expression of a constitutively active Akt
Ser/Thr kinase in 3T3-L1 adipocytes stimulates glucose uptake and glucose transporter 4
translocation. J Biol Chem 271: 31372–31378.
Kranendijk M, Struys EA, van Schaftingen E, Gibson KM, Kanhai WA, van der Knaap MS, Amiel J,
Buist NR, Das AM, de Klerk JB, et al. 2010. IDH2 mutations in patients with D-2-hydroxyglutaric
aciduria. Science 330: 336.
* Laplante M, Sabatini DM. 2012. mTOR signaling. Cold Spring Harb Perspect Biol 4: a011593.
Lee J, Kim HK, Han YM, Kim J. 2008. Pyruvate kinase isozyme type M2 (PKM2) interacts and
cooperates with Oct-4 in regulating transcription. Int J Biochem Cell Biol 40: 1043–1054.
Li B, Carey M, Workman JL. 2007. The role of chromatin during transcription. Cell 128: 707–719.
Lum JJ, Bui T, Gruber M, Gordan JD, DeBerardinis RJ, Covello KL, Simon MC, Thompson CB. 2007.
The transcription factor HIF-1α plays a critical role in the growth factor-dependent regulation of
both aerobic and anaerobic glycolysis. Genes Dev 21: 1037–1049.
Majmundar AJ, Wong WJ, Simon MC. 2010. Hypoxia-inducible factors and the response to hypoxic
stress. Mol Cell 40: 294–309.
Mardis ER, Ding L, Dooling DJ, Larson DE, McLellan MD, Chen K, Koboldt DC, Fulton RS,
Delehaunty KD, McGrath SD, et al. 2009. Recurring mutations found by sequencing an acute
myeloid leukemia genome. N Engl J Med 361: 1058–1066.
Mayer C, Zhao J, Yuan X, Grummt I. 2004. mTOR-dependent activation of the transcription factor
TIF-IA links rRNA synthesis to nutrient availability. Genes Dev 18: 423–434.
Metallo CM, Vander Heiden MG. 2010. Metabolism strikes back: Metabolic flux regulates cell
signaling. Genes Dev 24: 2717–2722.
* Morrison DK. 2012. MAP kinase pathways. Cold Spring Harb Perspect Biol 4: a011254.
Nicklin P, Bergman P, Zhang B, Triantafellow E, Wang H, Nyfeler B, Yang H, Hild M, Kung C,
Wilson C, et al. 2009. Bidirectional transport of amino acids regulates mTOR and autophagy. Cell
136: 521–534.
Oh WJ, Wu CC, Kim SJ, Facchinetti V, Julien LA, Finlan M, Roux PP, Su B, Jacinto E. 2010.
mTORC2 can associate with ribosomes to promote cotranslational phosphorylation and stability of
nascent Akt polypeptide. EMBO J 29: 3939–3951.
Osthus RC, Shim H, Kim S, Li Q, Reddy R, Mukherjee M, Xu Y, Wonsey D, Lee LA, Dang CV. 2000.
Deregulation of glucose transporter 1 and glycolytic gene expression by c-Myc. J Biol Chem 275:
21797–21800.
Papandreou I, Cairns RA, Fontana L, Lim AL, Denko NC. 2006. HIF-1 mediates adaptation to hypoxia
by actively downregulating mitochondrial oxygen consumption. Cell Metab 3: 187–197.
Porstmann T, Griffiths B, Chung YL, Delpuech O, Griffiths JR, Downward J, Schulze A. 2005.
PKB/Akt induces transcription of enzymes involved in cholesterol and fatty acid biosynthesis via
activation of SREBP. Oncogene 24: 6465–6481.
Rathmell JC, Fox CJ, Plas DR, Hammerman PS, Cinalli RM, Thompson CB. 2003. Akt-directed
glucose metabolism can prevent Bax conformation change and promote growth factor-independent
survival. Mol Cell Biol 23: 7315–7328.
Sancak Y, Thoreen CC, Peterson TR, Lindquist RA, Kang SA, Spooner E, Carr SA, Sabatini DM.
2007. PRAS40 is an insulin-regulated inhibitor of the mTORC1 protein kinase. Mol Cell 25: 903–
915.
Sancak Y, Peterson TR, Shaul YD, Lindquist RA, Thoreen CC, Bar-Peled L, Sabatini DM. 2008. The
Rag GTPases bind raptor and mediate amino acid signaling to mTORC1. Science 320: 1496–1501.
Sancak Y, Bar-Peled L, Zoncu R, Markhard AL, Nada S, Sabatini DM. 2010. Ragulator–Rag complex
targets mTORC1 to the lysosomal surface and is necessary for its activation by amino acids. Cell
141: 290–303.
Scholnick P, Lang D, Racker E. 1973. Regulatory mechanisms in carbohydrate metabolism. IX.
Stimulation of aerobic glycolysis by energy-linked ion transport and inhibition by dextran sulfate. J
Biol Chem 248: 5175–5182.
Selak MA, Armour SM, MacKenzie ED, Boulahbel H, Watson DG, Mansfield KD, Pan Y, Simon MC,
Thompson CB, Gottlieb E. 2005. Succinate links TCA cycle dysfunction to oncogenesis by
inhibiting HIF-α prolyl hydroxylase. Cancer Cell 7: 77–85.
Semenza GL, Roth PH, Fang HM, Wang GL. 1994. Transcriptional regulation of genes encoding
glycolytic enzymes by hypoxia-inducible factor 1. J Biol Chem 269: 23757–23763.
Sengupta S, Peterson TR, Sabatini DM. 2010. Regulation of the mTOR complex 1 pathway by
nutrients, growth factors, and stress. Mol Cell 40: 310–322.
Shatrov VA, Sumbayev VV, Zhou J, Brune B. 2003. Oxidized low-density lipoprotein (oxLDL)
triggers hypoxia-inducible factor-1α (HIF-1α) accumulation via redox-dependent mechanisms.
Blood 101: 4847–4849.
Shim H, Dolde C, Lewis BC, Wu CS, Dang G, Jungmann RA, Dalla-Favera R, Dang CV. 1997. c-Myc
transactivation of LDH-A: Implications for tumor metabolism and growth. Proc Natl Acad Sci 94:
6658–6663.
Taha C, Liu Z, Jin J, Al-Hasani H, Sonenberg N, Klip A. 1999. Opposite translational control of
GLUT1 and GLUT4 glucose transporter mRNAs in response to insulin. Role of mammalian target
of rapamycin, protein kinase b, and phosphatidylinositol 3-kinase in GLUT1 mRNA translation. J
Biol Chem 274: 33085–33091.
Tang H, Hornstein E, Stolovich M, Levy G, Livingstone M, Templeton D, Avruch J, Meyuhas O. 2001.
Amino acid–induced translation of TOP mRNAs is fully dependent on phosphatidylinositol 3-
kinase-mediated signaling, is partially inhibited by rapamycin, and is independent of S6K1 and
rpS6 phosphorylation. Mol Cell Biol 21: 8671–8683.
Tong X, Zhao F, Thompson CB. 2009. The molecular determinants of de novo nucleotide biosynthesis
in cancer cells. Curr Opin Genet Dev 19: 32–37.
Vander Heiden MG, Cantley LC, Thompson CB. 2009. Understanding the Warburg effect: The
metabolic requirements of cell proliferation. Science 324: 1029–1033.
Vander Heiden MG, Locasale JW, Swanson KD, Sharfi H, Heffron GJ, Amador-Noguez D, Christofk
HR, Wagner G, Rabinowitz JD, Asara JM, et al. 2010. Evidence for an alternative glycolytic
pathway in rapidly proliferating cells. Science 329: 1492–1499.
Wang X, Li W, Williams M, Terada N, Alessi DR, Proud CG. 2001. Regulation of elongation factor 2
kinase by p90 (RSK1) and p70 S6 kinase. EMBO J 20: 4370–4379.
Ward PS, Patel J, Wise DR, Abdel-Wahab O, Bennett BD, Coller HA, Cross JR, Fantin VR, Hedvat
CV, Perl AE, et al. 2010. The common feature of leukemia-associated IDH1 and IDH2 mutations is
a neomorphic enzyme activity converting α-ketoglutarate to 2-hydroxyglutarate. Cancer Cell 17:
225–234.
Ward PS, Cross JR, Lu C, Weigert O, Abel-Wahab O, Levine RL, Weinstock DM, Sharp KA,
Thompson CB. 2011. Identification of additional IDH mutations associated with oncometabolite
R(−)-2-hydroxyglutarate production. Oncogene doi: 10.1038/onc.2011.416.
Wellen KE, Thompson CB. 2010. Cellular metabolic stress: Considering how cells respond to nutrient
excess. Mol Cell 40: 323–332.
Wellen KE, Hatzivassiliou G, Sachdeva UM, Bui TV, Cross JR, Thompson CB. 2009. ATP-citrate
lyase links cellular metabolism to histone acetylation. Science 324: 1076–1080.
Wellen KE, Lu C, Mancuso A, Lemons JM, Ryczko M, Dennis JW, Rabinowitz JD, Coller HA,
Thompson CB. 2010. The hexosamine biosynthetic pathway couples growth factor-induced
glutamine uptake to glucose metabolism. Genes Dev 24: 2784–2799.
Wise DR, Thompson CB. 2010. Glutamine addiction: A new therapeutic target in cancer. Trends
Biochem Sci 35: 427–433.
Wise DR, DeBerardinis RJ, Mancuso A, Sayed N, Zhang XY, Pfeiffer HK, Nissim I, Daikhin E,
Yudkoff M, McMahon SB, et al. 2008. Myc regulates a transcriptional program that stimulates
mitochondrial glutaminolysis and leads to glutamine addiction. Proc Natl Acad Sci 105: 18782–
18787.
Wu L, Derynck R. 2009. Essential role of TGF-β signaling in glucose-induced cell hypertrophy. Dev
Cell 17: 35–48.
Yan H, Parsons DW, Jin G, McLendon R, Rasheed BA, Yuan W, Kos I, Batinic-Haberle I, Jones S,
Riggins GJ, et al. 2009. IDH1 and IDH2 mutations in gliomas. N Engl J Med 360: 765–773.
Yuneva M, Zamboni N, Oefner P, Sachidanandam R, Lazebnik Y. 2007. Deficiency in glutamine but
not glucose induces MYC-dependent apoptosis in human cells. J Cell Biol 178: 93–105.
Zhong H, Chiles K, Feldser D, Laughner E, Hanrahan C, Georgescu MM, Simons JW, Semenza GL.
2000. Modulation of hypoxia-inducible factor 1α expression by the epidermal growth
factor/phosphatidylinositol 3-kinase/PTEN/AKT/FRAP pathway in human prostate cancer cells:
Implications for tumor angiogenesis and therapeutics. Cancer Res 60: 1541–1545.
Zinzalla V, Stracka D, Oppliger W, Hall MN. 2011. Activation of mTORC2 by association with the
ribosome. Cell 144: 757–768.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a006783
CHAPTER 8
SUMMARY
Outline
1 Introduction
2 The migration machinery
3 Migration signaling networks
4 Biased excitable biochemical networks in chemotaxis
5 Concluding remarks
References
1 INTRODUCTION
Cell migration plays a pivotal role in a wide variety of phenomena
throughout phylogeny (Trinkaus 1969; Ridley et al. 2003; Lee et al. 2005;
Cai et al. 2010). In the amoeba Dictyostelium, it functions in nutrient seeking,
cell–cell aggregation, and the morphogenesis of multicellular structures. In
metazoans, cells migrate both throughout embryogenesis and in the adult. In
early developmental events, such as gastrulation or dorsal closure, large cell
sheets migrate and fold; at later stages, precursor cells that reside in the
neural crest, somites, brain ventricles, and other stem cell regions leave
epithelial sheets and migrate to their target destinations. In the adult, cell
migrations are critical for immune cell trafficking, wound healing, and stem
cell homing, among other processes. A closely related phenomenon is the
directed growth of specialized cellular extensions (e.g., yeast mating
“shmoos” [Slessareva and Dohlman 2006], pollen tubes [Takeuchi and
Higashiyama 2011], and the dendrites, axons, and spines of neurons [Tada
and Sheng 2006; Geraldo and Gordon-Weeks 2009]).
Most migrating cells and cellular extensions have an internal compass that
enables them to sense and move along gradients of soluble attractants and
repellents, a process referred to as chemotaxis (Devreotes and Janetopoulos
2003). Many chemoattractants act through G-protein-coupled receptors
(GPCRs). Examples include cAMP acting on cAMP receptors in
Dictyostelium and chemokines such as SDF1 acting on chemokine receptors
in metazoans. Growth factors acting on receptor tyrosine kinases and
cytokines such as transforming growth factor β (TGFβ) also function as
chemoattractants. Like migrating cells, the growth of axons can be guided by
a series of extracellular protein attractants and repellents (e.g., nephrins).
Cells can also be guided by gradients of immobilized signaling molecules
(haplotaxis), substrate rigidity (durotaxis), electric fields (galvanotaxis), and
shear force (mechanotaxis). Importantly, many diseases involve defective or
unregulated cell migration or protrusive growth (Ridley et al. 2003). For
example, tumor invasion and metastasis occur as a consequence of the
movement of both individual cells and large collectives (Friedl and Gilmour
2009; Friedl and Alexander 2011), arthritis and asthma result from excessive
migration of inflammatory cells (Montoya et al. 2002; Vicente-Manzanares et
al. 2002), and several cognitive disorders are accompanied by abnormal
neuronal extensions (Newey et al. 2005; van Galen and Ramakers 2005).
Cell migration requires coordination of cytoskeletal dynamics and
reorganization, cell adhesion, and signal transduction, and takes a variety of
forms (see Box 1) (Lauffenburger and Horwitz 1996; Mitchison and Cramer
1996; Ridley et al. 2003). Here, we first examine the machinery that drives
migration—the actin cytoskeleton, cell adhesions, and their regulators. We
then discuss signaling networks that control the migration machinery, starting
with those closest to the cytoskeleton then adding upstream components.
Finally, we address how chemotactic cues regulate motility. There are, of
course, other kinds of motility, such as sperm and cilial motility, but they use
microtubule-based mechanisms and are not addressed here.
2.2 Adhesion
Adhesion to the substrate is common to most migrating cells. Integrin-
mediated adhesions are the best studied (Hynes 2002). The integrins are a
large family of heterodimeric transmembrane receptors that link to actin via a
specialized set of molecules that include talin, vinculin, and α-actinin. The
adhesions in which these components reside are large assemblies containing
>150 different molecules that mediate intracellular signaling in addition to
adhesion to proteins in the extracellular matrix, such as fibronectin and
laminin (Zaidel-Bar et al. 2007; Parsons et al. 2010; Zaidel-Bar and Geiger
2010). The affinity of integrins is regulated by the binding of talin and
kindlin, cytoplasmic proteins that bind directly to the cytoplasmic domain of
the integrin β subunit, and also by phosphatidylinositol 4,5-bisphosphate
(PIP2) and other adhesion-associated molecules (Moser et al. 2009; Shattil et
al. 2010).
Adhesions serve both as traction points and as signaling centers during
cell migration (Parsons et al. 2010). As traction points, they transmit forces to
the substrate so that actin polymerization causes protrusion at the cell front.
These traction points are released at the cell rear as it retracts and the cell
moves forward. Although this release is efficient in some cells and substrates,
it is not in others and can be rate limiting for migration (Lauffenburger and
Horwitz 1996). Thus, there is an optimum strength of attachment that allows
sufficient adhesion for traction at the cell front and yet allows for efficient
release at the rear (Palecek et al. 1997). As signaling centers, adhesions in
protrusions regulate actin polymerization and myosin II activity through Rho-
family GTPases (Parsons et al. 2010). Although adhesions can vary
considerably in size, location, and presumably function, they have not yet
been classified clearly and meaningfully based on differences in composition
and function. Adhesions in vivo tend to be small and dynamic in migrating
cells; however, highly elongated adhesions have also been observed
(Harunaga and Yamada 2011; Kubow and Horwitz 2011).
The cytoskeleton in turn regulates adhesions via an incompletely
understood feedback loop involving actin polymerization and myosin-II-
mediated contraction (Fig. 2). Nascent adhesions form in the region of
dendritic actin, and their formation is coupled to actin polymerization
(Alexandrova et al. 2008; Choi et al. 2008). At the interface of dendritic actin
in the lamellipodium and the actin bundles in the adjacent lamellum,
adhesions elongate along actin filament bundles (Small et al. 2002; Choi et
al. 2008; Geiger and Yamada 2011; Oakes et al. 2012). The fraction of
adhesions that grow, as well as the extent of maturation, is determined at least
in part by myosin II activity (Vicente-Manzanares et al. 2009; Oakes et al.
2012). Proteases such as calpain, a calcium-activated protease, mediate
adhesion disassembly by acting on adhesion proteins such as talin, which link
actin and integrins (Franco et al. 2004; Chan et al. 2010; Cortesio et al.
2011). The repeated direct contact between microtubule tips and adhesions
and endocytosis of integrins driven by the GTPase dynamin also contribute to
disassembly (Kaverina et al. 1999; Broussard et al. 2008; Ezratty et al. 2009;
Gerisch et al. 2011).
Figure 2. Adhesions serve as contact points and signaling centers. Integrin-based adhesions are large,
complex assemblies that link the substratum to actin and generate signals that regulate Rho GTPases
and cell migration. The structural linkage to actin is thought to be mediated by talin, vinculin, and
perhaps α-actinin. The signaling is mediated by adhesion-associated complexes. The paxillin/FAK
module and its link to some Rac GEFs and Rho GEFs is shown as an example. The Arp2/3 complex
and myosin II, whose activity is regulated by Rho and Rac, are also shown.
2.3 Polarization
The presence of a distinct front and rear is a key feature of cell migration.
Some cells can polarize spontaneously and migrate in a directionally
persistent manner. The machinery that establishes polarity is incompletely
understood but microtubules, vesicle cycling, and actomyosin filaments
appear to be the drivers. In epithelial cells and astrocytes, polarity is
established through a signaling pathway involving Cdc42, Par3/6, and
atypical protein kinase C (aPKC) that targets microtubules (Etienne-
Manneville and Hall 2002; Etienne-Manneville et al. 2005). This pathway
orients the microtubule-organizing center (MTOC) and Golgi apparatus (Ch.
9 [McCaffrey and Macara 2012]). In fibroblasts, activated myosin II creates a
region of actomyosin filament bundles that terminate in adhesions that do not
contain guanine nucleotide exchange factors (GEFs) and therefore do not
support Rac or Cdc42 signaling and actin polymerization (Vicente-
Manzanares et al. 2008, 2011). This region becomes the rear and sides, and
zones of active Rac generate protrusions that elongate the cell to form the
front; the actomyosin system appears to set up the initial polarity, which is
then refined by the microtubule system (Vicente-Manzanares et al. 2008).
Figure 4. A portion of the Dictyostelium migration signaling network involving PIP3 and TorC2.
Colored blocks delineate modules. The overlapping of blocks indicates that some components belong to
several modules. CARE, cystic AMP receptors.
The LEGI model is a useful conceptual device that allows one to predict
the response to any combination of applied temporal and spatial stimuli, but
further studies are needed to define the underlying biochemical events and to
link the model to cell migration. First, the excitatory process likely
corresponds to G-protein activation. When cells are exposed to
chemoattractant, the G-protein subunits dissociate within a few seconds and
all of the biochemical responses in the network are triggered. During the next
several minutes, the responses gradually subside even though the G protein
does not reassociate. The mechanism that offsets the activity of the G-protein
and causes the responses to subside remains to be determined. Second, LEGI
schemes can account for all of the behavior of immobilized cells but fail to
explain migration or polarity. However, the output of LEGI could enhance
excitability at the front and suppress it at the rear (Xiong et al. 2010). This
would ensure that the system responds to the steepness of a gradient but is
independent of its midpoint concentration. LEGI-BEN schemes are capable
of extraordinary sensitivity, and computer simulations show that this model
can produce realistic temporal and spatial chemotactic responses.
5 CONCLUDING REMARKS
Many of the principles described in this chapter appear to be general and
apply to cellular behaviors analogous to migration. For example, dendritic
spines, small extensions along dendrites of neurons in the central nervous
system, contain a highly organized postsynaptic density that receives
excitatory signals. Like protrusions in migrating cells, dendritic spines are
highly dynamic, they undergo complex morphologic changes, and they
contain a highly organized adhesion associated with the postsynaptic density.
Actin polymerization and actomyosin activity play a major role in spine and
postsynaptic density (PSD) organization (Oertner and Matus 2005; Hodges et
al. 2011). Rho-family GTPases have emerged as major regulators of spine
organization and dynamics and are implicated in human cognitive diseases.
Indeed, mutations in regulators of Rho-family GTPases are implicated in
spine-related diseases including autism, schizophrenia, and nonsyndromic
mental retardation. With respect to the latter, α-Pix (a Rac GEF), PAK3 (a
Rac/Cd42 effector), and oligphrenin 1 (a Rho GAP) are all associated with
nonsyndromic mental retardation in humans—a disease characterized by
spine defects (van Galen and Ramakers 2005). Pix and PAK are localized to
dendritic spines via GIT1 and thought to spatially restrict their formation
(Zhang et al. 2005). Asymmetric localization of PIP3 is observed in many
migrating cells. Oncogenic mutations leading to overproduction of PIP3 are
typically assumed to increase growth rates; but it is likely that many of the
cancer-causing effects can also be attributed to alterations in the cytoskeleton
(Kim et al. 2011). Distortion of cell migration signaling networks plays a
critical role in migration-related diseases such as invasive and metastatic
cancer (Ch. 21 [Sever and Brugge 2014]). The plethora of signaling pathways
that converge on Rho GTPases means there is a very large potential set of
loci for the misregulation of migration. It also suggests that drugs directed
against any particular pathway may not be effective for long given the
selection that occurs in the tumor environment. However, the convergence on
Rho-family GTPases and the limited migration machinery on which it acts
hold promise for therapeutic strategies targeting these GTPases and their
downstream effectors or diagnostic routes to identifying cells with invasive
potential.
Another emerging area of research is the study of migration in 3D. Until
recently, most migration studies focused on migration on planar substrates
using integrin-mediated adhesion. Analogous mechanisms are probably used
by cells migrating in 3D, or using other receptors (e.g., in the central nervous
system). However, different pathways might play more prominent roles in
each case. For example, in 3D, cell protrusions, adhesion, and cell
morphology all appear to differ from that generally seen on rigid planar 2D
substrates (Even-Ram and Yamada 2005; Provenzano et al. 2009; Friedl and
Wolf 2010; Sanz-Moreno and Marshall 2010). In 3D, the cells are more
elongated and possess narrower protrusions and smaller adhesions (Harunaga
and Yamada 2011). Moreover, there is evidence that different signaling
pathways are indeed involved. For example, depleting paxillin produces a
mesenchymal phenotype in 3D environments, whereas depleting the paxillin
relative Hic5 produces an amoeboid morphology (Deakin and Turner 2008,
2011). The particular signaling pathway used seems to depend on the cellular
microenvironment, which can differ between normal and tumor cells.
Research into cell migration has clearly made enormous progress. The
basic machines that drive migration have been described, and many of the
pathways that regulate them have been identified. However, we have only
scratched the surface and much remains to be understood. The interactions
and regulation of the complex signaling networks that orchestrate migration
and the mechanism by which extracellular forces affect these networks are
not understood. Furthermore, new modes of migration are being uncovered,
including blebbing-mediated migration and the newly described lobopodia
migration (Petrie et al. 2012). In addition, migration in complex in vivo
environments differs from that seen on rigid planar substrates and its study
presents unexpected challenges. Finally, integrative, quantitative models of
migration that conjoin the plethora of regulatory networks are only now
beginning to be developed.
REFERENCES
*Reference is in this book.
Alexandrova AY, Arnold K, Schaub S, Vasiliev JM, Meister JJ, Bershadsky AD, Verkhovsky AB.
2008. Comparative dynamics of retrograde actin flow and focal adhesions: Formation of nascent
adhesions triggers transition from fast to slow flow. PLoS ONE 3: e3234.
Arai Y, Shibata T, Matsuoka S, Sato MJ, Yanagida T, Ueda M. 2010. Self-organization of the
phosphatidylinositol lipids signaling system for random cell migration. Proc Natl Acad Sci 107:
12399–12404.
Bamburg JR, Bernstein BW. 2008. ADF/cofilin. Curr Biol 18: R273–R275.
Bear JE, Gertler FB. 2009. Ena/VASP: Towards resolving a pointed controversy at the barbed end. J
Cell Sci 122: 1947–1953.
Bershadsky AD, Balaban NQ, Geiger B. 2003. Adhesion-dependent cell mechanosensitivity. Annu Rev
Cell Dev Biol 19: 677–695.
Bokoch GM. 1995. Chemoattractant signaling and leukocyte activation. Blood 86: 1649–1660.
Bokoch GM. 2003. Biology of the p21-activated kinases. Annu Rev Biochem 72: 743–781.
Bretschneider T, Anderson K, Ecke M, Muller-Taubenberger A, Schroth-Diez B, Ishikawa-Ankerhold
HC, Gerisch G. 2009. The three-dimensional dynamics of actin waves, a model of cytoskeletal self-
organization. Biophys J 96: 2888–2900.
Broussard JA, Webb DJ, Kaverina I. 2008. Asymmetric focal adhesion disassembly in motile cells.
Curr Opin Cell Biol 20: 85–90.
Brown MC, Turner CE. 2004. Paxillin: Adapting to change. Physiol Rev 84: 1315–1339.
Brown CM, Hebert B, Kolin DL, Zareno J, Whitmore L, Horwitz AR, Wiseman PW. 2006. Probing the
integrin-actin linkage using high-resolution protein velocity mapping. J Cell Sci 119: 5204–5214.
Bugyi B, Carlier MF. 2010. Control of actin filament treadmilling in cell motility. Annu Rev Biophys
39: 449–470.
Cai H, Das S, Kamimura Y, Comer FI, Parent DA, Devreotes PN. 2010. Ras-mediated activation and
inactivation of the TorC2-PKB pathway are critical for chemotaxis. J Cell Biol 190: 233–245.
Case LB, Waterman CM. 2011. Adhesive F-actin waves: A novel integrin-mediated adhesion complex
coupled to ventral actin polymerization. PLoS ONE 6: e26631.
Chadborn NH, Ahmed AI, Holt MR, Prinjha R, Dunn GA, Jones GE, Eickholt BJ. 2006. PTEN couples
Sema3A signalling to growth cone collapse. J Cell Sci 119: 951–957.
Chan KT, Bennin DA, Huttenlocher A. 2010. Regulation of adhesion dynamics by calpain-mediated
proteolysis of focal adhesion kinase (FAK). J Biol Chem 285: 11418–11426.
Charest PG, Shen Z, Lakoduk A, Sasaki AT, Briggs SP, Firtel RA. 2010. A Ras signaling complex
controls the RasC-TORC2 pathway and directed cell migration. Dev Cell 18: 737–749.
Charras G, Paluch E. 2008. Blebs lead the way: How to migrate without lamellipodia. Nat Rev Mol Cell
Biol 9: 730–736.
Chen M-Y, Long Y, Devreotes PN. 1997. A novel cytosolic regulator, Pianissimo, is required for
chemoattractant receptor and G protein-mediated activation of the twelve transmembrane domain
adenylyl cyclase in Dictyostelium. Genes Dev 11: 3218–3231.
Chen L, Janetopoulos C, Huang YE, Iijima M, Borleis J, Devreotes PN. 2003. Two phases of actin
polymerization display different dependences on PI(3,4,5)P3 accumulation and have unique roles
during chemotaxis. Mol Biol Cell 14: 5028–5037.
Chen L, Iijima M, Tang M, Landree MA, Huang YE, Xiong Y, Iglesias PA, Devreotes PN. 2007. PLA2
and PI3K/PTEN pathways act in parallel to mediate chemotaxis. Dev Cell 12: 603–614.
Chen L, Vicente-Manzanares M, Potvin-Trottier L, Wiseman PW, Horwitz AR. 2012. The integrin-
ligand interaction regulates adhesion and migration through a molecular clutch. PLoS ONE 7:
e40202.
Chesarone MA, DuPage AG, Goode BL. 2010. Unleashing formins to remodel the actin and
microtubule cytoskeletons. Nat Rev Mol Cell Biol 11: 62–74.
Choi CK, Vicente-Manzanares M, Zareno J, Whitmore LA, Mogilner A, Horwitz AR. 2008. Actin and
α-actinin orchestrate the assembly and maturation of nascent adhesions in a myosin II motor-
independent manner. Nat Cell Biol 10: 1039–1050.
Choi CK, Zareno J, Digman MA, Gratton E, Horwitz AR. 2011. Cross-correlated fluctuation analysis
reveals phosphorylation-regulated paxillin-FAK complexes in nascent adhesions. Biophys J 100:
583–592.
Clarke M, Muller-Taubenberger A, Anderson KI, Engel U, Gerisch G. 2006. Mechanically induced
actin-mediated rocketing of phagosomes. Mol Biol Cell 17: 4866–4875.
Cohen DM, Kutscher B, Chen H, Murphy DB, Craig SW. 2006. A conformational switch in vinculin
drives formation and dynamics of a talin-vinculin complex at focal adhesions. J Biol Chem 281:
16006–16015.
Comer FI, Parent CA. 2007. Phosphoinositides specify polarity during epithelial organ development.
Cell 128: 239–240.
Condeelis J. 2001. How is actin polymerization nucleated in vivo? Trends Cell Biol 11: 288–293.
Cortesio CL, Boateng LR, Piazza TM, Bennin DA, Huttenlocher A. 2011. Calpain-mediated
proteolysis of paxillin negatively regulates focal adhesion dynamics and cell migration. J Biol
Chem 286: 9998–10006.
Deakin NO, Turner CE. 2008. Paxillin comes of age. J Cell Sci 121: 2435–2444.
Deakin NO, Turner CE. 2011. Distinct roles for paxillin and Hic-5 in regulating breast cancer cell
morphology, invasion, and metastasis. Mol Biol Cell 22: 327–341.
del Rio A, Perez-Jimenez R, Liu R, Roca-Cusachs P, Fernandez JM, Sheetz MP. 2009. Stretching
single talin rod molecules activates vinculin binding. Science 323: 638–641.
DerMardirossian C, Bokoch GM. 2005. GDIs: Central regulatory molecules in Rho GTPase activation.
Trends Cell Biol 15: 356–363.
Devreotes P, Janetopoulos C. 2003. Eukaryotic chemotaxis: Distinctions between directional sensing
and polarization. J Biol Chem 278: 20445–20448.
Devreotes PN, Zigmond SH. 1988. Chemotaxis in eucaryotic cells: A focus on leukocytes and
Dictyostelium. Ann Rev Cell Biol 4: 649–686.
Döbereiner HG, Dubin-Thaler BJ, Hofman JM, Xenias HS, Sims TN, Giannone G, Dustin ML,
Wiggins CH, Sheetz MP. 2006. Lateral membrane waves constitute a universal dynamic pattern of
motile cells. Phys Rev Lett 97: 038102.
Dumstrei K, Mennecke R, Raz E. 2004. Signaling pathways controlling primordial germ cell migration
in zebrafish. J Cell Sci 117: 4787–4795.
Etienne-Manneville S. 2004. Cdc42—The centre of polarity. J Cell Sci 117: 1291–1300.
Etienne-Manneville S, Hall A. 2002. Rho GTPases in cell biology. Nature 420: 629–635.
Etienne-Manneville S, Hall A. 2003. Cell polarity: Par6, aPKC and cytoskeletal crosstalk. Curr Opin
Cell Biol 15: 67–72.
Etienne-Manneville S, Manneville JB, Nicholls S, Ferenczi MA, Hall A. 2005. Cdc42 and Par6-PKCζ
regulate the spatially localized association of Dlg1 and APC to control cell polarization. J Cell Biol
170: 895–901.
Evans JH, Falke JJ. 2007. Ca2+ influx is an essential component of the positive-feedback loop that
maintains leading-edge structure and activity in macrophages. Proc Natl Acad Sci 104: 16176–
16181.
Even-Ram S, Yamada KM. 2005. Cell migration in 3D matrix. Curr Opin Cell Biol 17: 524–532.
Ezratty EJ, Partridge MA, Gundersen GG. 2005. Microtubule-induced focal adhesion disassembly is
mediated by dynamin and focal adhesion kinase. Nat Cell Biol 7: 581–590.
Ezratty EJ, Bertaux C, Marcantonio EE, Gundersen GG. 2009. Clathrin mediates integrin endocytosis
for focal adhesion disassembly in migrating cells. J Cell Biol 187: 733–747.
Fackler OT, Grosse R. 2008. Cell motility through plasma membrane blebbing. J Cell Biol 181: 879–
884.
Ferguson GJ, Milne L, Kulkarni S, Sasaki T, Walker S, Andrews S, Crabbe T, Finan P, Jones G,
Jackson S, et al. 2007. PI(3)Kγ has an important context-dependent role in neutrophil chemokinesis.
Nat Cell Biol 9: 86–91.
Frame MC, Patel H, Serrels B, Lietha D, Eck MJ. 2010. The FERM domain: Organizing the structure
and function of FAK. Nat Rev Mol Cell Biol 11: 802–814.
Franco SJ, Rodgers MA, Perrin BJ, Han J, Bennin DA, Critchley DR, Huttenlocher A. 2004. Calpain-
mediated proteolysis of talin regulates adhesion dynamics. Nat Cell Biol 6: 977–983.
Friedl P, Alexander S. 2011. Cancer invasion and the microenvironment: Plasticity and reciprocity. Cell
147: 992–1009.
Friedl P, Gilmour D. 2009. Collective cell migration in morphogenesis, regeneration and cancer. Nat
Rev Mol Cell Biol 10: 445–457.
Friedl P, Wolf K. 2010. Plasticity of cell migration: A multiscale tuning model. J Cell Biol 188: 11–19.
Funamoto S, Meili R, Lee S, Parry L, Firtel RA. 2002. Spatial and temporal regulation of 3-
phosphoinositides by PI 3-kinase and PTEN mediates chemotaxis. Cell 109: 611–623.
Garcia-Mata R, Boulter E, Burridge K. 2011. The “invisible hand”: Regulation of Rho GTPases by Rho
GDIs. Nat Rev Mol Cell Biol 12: 493–504.
Geiger B, Yamada KM. 2011. Molecular architecture and function of matrix adhesions. Cold Spring
Harb Perspect Biol 3: a005033.
Geraldo S, Gordon-Weeks PR. 2009. Cytoskeletal dynamics in growth-cone steering. J Cell Sci 122:
3595–3604.
Gerisch G. 2010. Self-organizing actin waves that simulate phagocytic cup structures. PMC Biophys 3:
7.
Gerisch G, Bretschneider T, Müller-Taubenberger A, Simmeth E, Ecke M, Diez S, Anderson K. 2004.
Mobile actin clusters and traveling waves in cells recovering from actin depolymerization. Biophys
J 87: 3493–3503.
Gerisch G, Ecke M, Wischnewski D, Schroth-Diez B. 2011. Different modes of state transitions
determine pattern in the Phosphatidylinositide-Actin system. BMC Cell Biol 12: 42.
Giannone G, Dubin-Thaler BJ, Döbereiner HG, Kieffer N, Bresnick AR, Sheetz MP. 2004. Periodic
lamellipodial contractions correlate with rearward actin waves. Cell 116: 431–443.
Goode BL, Eck MJ. 2007. Mechanism and function of formins in the control of actin assembly. Annu
Rev Biochem 76: 593–627.
Grande-Garcia A, Echarri A, Del Pozo MA. 2005. Integrin regulation of membrane domain trafficking
and Rac targeting. Biochem Soc Trans 33: 609–613.
Grashoff C, Hoffman BD, Brenner MD, Zhou R, Parsons M, Yang MT, McLean MA, Sligar SG, Chen
CS, Ha T, et al. 2010. Measuring mechanical tension across vinculin reveals regulation of focal
adhesion dynamics. Nature 466: 263–266.
Harunaga JS, Yamada KM. 2011. Cell-matrix adhesions in 3D. Matrix Biol 30: 363–368.
Heasman SJ, Ridley AJ. 2008. Mammalian Rho GTPases: New insights into their functions from in
vivo studies. Nat Rev Mol Cell Biol 9: 690–701.
Hecht I, Skoge ML, Charest PG, Ben-Jacob E, Firtel RA, Loomis WF, Levine H, Rappel WJ. 2011.
Activated membrane patches guide chemotactic cell motility. PLoS Comput Biol 7: e1002044.
* Hemmings BA, Restuccia DF. 2012. PI3K-PKB/Akt pathway. Cold Spring Harb Perspect Biol 4:
a011189.
Hodges JL, Newell-Litwa K, Asmussen H, Vicente-Manzanares M, Horwitz AR. 2011. Myosin IIb
activity and phosphorylation status determines dendritic spine and post-synaptic density
morphology. PLoS ONE 6: e24149.
Hoefen RJ, Berk BC. 2006. The multifunctional GIT family of proteins. J Cell Sci 119: 1469–1475.
Hoeller O, Kay RR. 2007. Chemotaxis in the absence of PIP3 gradients. Curr Biol 17: 813–817.
Hu K, Ji L, Applegate KT, Danuser G, Waterman-Storer CM. 2007. Differential transmission of actin
motion within focal adhesions. Science 315: 111–115.
Huang YE, Iijima M, Parent CA, Funamoto S, Firtel RA, Devreotes P. 2003. Receptor-mediated
regulation of PI3Ks confines PI(3,4,5)P3 to the leading edge of chemotaxing cells. Mol Biol Cell
14: 1913–1922.
Huttenlocher A, Horwitz AR. 2011. Integrins in cell migration. Cold Spring Harb Perspect Biol 3:
a005074.
Huttenlocher A, Palecek SP, Lu Q, Zhang W, Mellgren RL, Lauffenburger DA, Ginsberg MH, Horwitz
AF. 1997. Regulation of cell migration by the calcium-dependent protease calpain. J Biol Chem
272: 32719–32722.
Hynes RO. 2002. Integrins: Bidirectional, allosteric signaling machines. Cell 110: 673–687.
Iglesias PA, Devreotes PN. 2012. Biased excitable networks: How cells direct motion in response to
gradients. Curr Opin Cell Biol 757: 451–468.
Iglesias PA, Levchenko A. 2002. Modeling the cell’s guidance system. Sci STKE 2002: re12.
Iijima M, Devreotes PN. 2002. Tumor suppressor PTEN mediates sensing of chemoattractant gradients.
Cell 109: 599–610.
Inoue T, Meyer T. 2008. Synthetic activation of endogenous PI3K and Rac identifies an AND-gate
switch for cell polarization and migration. PLoS ONE 3: e3068.
Insall RH, Machesky LM. 2009. Actin dynamics at the leading edge: From simple machinery to
complex networks. Dev Cell 17: 310–322.
Janetopoulos C, Devreotes PN. 2006. Phosphoinositide signaling plays a key role in cytokinesis. J Cell
Biol 174: 485–490.
Janetopoulos C, Ma L, Iglesias PA, Devreotes PN. 2004. Chemoattractant-induced temporal and spatial
PI(3,4,5)P3 accumulation is controlled by a local excitation, global inhibition mechanism. Proc
Natl Acad Sci 101: 8951–8956.
Janetopoulos C, Borleis J, Vazquez F, Iijima M, Devreotes P. 2005. Temporal and spatial regulation of
phosphoinositide signaling mediates cytokinesis. Dev Cell 8: 467–477.
Jay DG. 2000. The clutch hypothesis revisited: Ascribing the roles of actin-associated proteins in
filopodial protrusion in the nerve growth cone. J Neurobiol 44: 114–125.
Kae H, Lim CJ, Spiegelman GB, Weeks G. 2004. Chemoattractants-induced Ras activation during
Dictyostelium aggregation. EMBO Rep 5: 602–606.
Kamimura Y, Xiong Y, Iglesias PA, Hoeller O, Bolourani P, Devreotes PN. 2008. PIP3-independent
activation of TorC2 and PKB at the cell’s leading edge mediates chemotaxis. Curr Biol 18: 1034–
1043.
Kaverina I, Krylyshkina O, Small JV. 1999. Microtubule targeting of substrate contacts promotes their
relaxation and dissociation. J Cell Biol 146: 1033–1044.
Kim EK, Yun SJ, HA JM, Kim YW, Jin IH, Yun J, Shin HK, Song SH, Kim JH, Lee JS, et al. 2011.
Selective activation of Akt1 by mammalian target of rapamycin complex 2 regulates cancer cell
migration, invasion, and metastasis. Oncogene 30: 2954–2963.
Krause M, Dent EW, Bear JE, Loureiro JJ, Gertler FB. 2003. Ena/VASP proteins: Regulators of the
actin cytoskeleton and cell migration. Annu Rev Cell Dev Biol 19: 541–564.
Kraynov VS, Chamberlain C, Bokoch GM, Schwartz MA, Slabaugh S, Hahn KM. 2000. Localized Rac
activation dynamics visualized in living cells. Science 290: 333–337.
Kubow KE, Horwitz AR. 2011. Reducing background fluorescence reveals adhesions in 3D matrices.
Nat Cell Biol 13: 3–5.
Lacalle RA, Gómez-Moutón C, Barber DF, Jiménez-Baranda S, Mira E, Martínez-A C, Carrera AC,
Mañes S. 2004. PTEN regulates motility but not directionality during leukocyte chemotaxis. J Cell
Sci 117: 6207–6215.
* Laplante M, Sabatini DM. 2012. mTOR signaling. Cold Spring Harb Perspect Biol 4: a011593.
Lauffenburger DA, Horwitz AF. 1996. Cell migration: A physically integrated molecular process. Cell
84: 359–369.
Lee S, Comer FI, Sasaki A, McLeod IX, Duong Y, Okumura K, Yates JR 3rd, Parent CA, Firtel RA.
2005. TOR complex 2 integrates cell movement during chemotaxis and signal relay in
Dictyostelium. Mol Biol Cell 16: 4572–4583.
Legate KR, Montañez E, Kudlacek O, Fässler R. 2006. ILK, PINCH and parvin: The tIPP of integrin
signalling. Nat Rev Mol Cell Biol 7: 20–31.
Levine H, Kessler DA, Rappel WJ. 2006. Directional sensing in eukaryotic chemotaxis: A balanced
inactivation model. Proc Natl Acad Sci 103: 9761–9766.
Linder S, Wiesner C, Himmel M. 2011. Degrading devices: Invadosomes in proteolytic cell invasion.
Annu Rev Cell Dev Biol 27: 185–211.
Liu L, Parent CA. 2011. TOR kinase complexes and cell migration. J Cell Biol 194: 815–824.
Liu L, Das S, Losert W, Parent CA. 2010. mTORC2 regulates neutrophil chemotaxis in a cAMP- and
RhoA-dependent fashion. Dev Cell 19: 845–857.
Machacek M, Hodgson L, Welch C, Elliott H, Pertz O, Nalbant P, Abell A, Johnson GL, Hahn KM,
Danuser G. 2009. Coordination of Rho GTPase activities during cell protrusion. Nature 461: 99–
103.
* McCaffrey LM, Macara IG. 2012. Signaling pathways in cell polarity. Cold Spring Harb Perspect
Biol 4: a009654.
Meili R, Ellsworth C, Lee S, Reddy TB, Ma H, Firtel RA. 1999. Chemoattractant-mediated transient
activation and membrane localization of Akt/PKB is required for efficient chemotaxis to cAMP in
Dictyostelium. EMBO J 18: 2092–2105.
Mitchison TJ, Cramer LP. 1996. Actin-based cell motility and cell locomotion. Cell 84: 371–379.
Mitchison T, Kirschner M. 1988. Cytoskeletal dynamics and nerve growth. Neuron 1: 761–772.
Mitra SK, Hanson DA, Schlaepfer DD. 2005. Focal adhesion kinase: In command and control of cell
motility. Nat Rev Mol Cell Biol 6: 56–68.
Mondal S, Subramanian KK, Sakai J, Bajrami B, Luo HR. 2012. Phosphoinositide lipid phosphatase
SHIP1 and PTEN coordinate to regulate cell migration and adhesion. Mol Biol Cell 23: 1219–1230.
Montoya MC, Sancho D, Vicente-Manzanares M, Sánchez-Madrid F. 2002. Cell adhesion and polarity
during immune interactions. Immunol Rev 186: 68–82.
Moser M, Legate KR, Zent R, Fässler R. 2009. The tail of integrins, talin, and kindlins. Science 324:
895–899.
Nalbant P, Hodgson L, Kraynov V, Toutchkine A, Hahn KM. 2004. Activation of endogenous Cdc42
visualized in living cells. Science 305: 1615–1619.
Newey SE, Velamoor V, Govek EE, Van Aelst L. 2005. Rho GTPases, dendritic structure, and mental
retardation. J Neurobiol 64: 58–74.
Nishio M, Watanabe K, Sasaki J, Taya C, Takasuga S, Iizuka R, Balla T, Yamazaki M, Watanabe H,
Itoh R, et al. 2007. Control of cell polarity and motility by the PtdIns(3,4,5)P3 phosphatase SHIP1.
Nat Cell Biol 9: 36–44.
Oakes PW, Beckham Y, Stricker J, Gardel ML. 2012. Tension is required but not sufficient for focal
adhesion maturation without a stress fiber template. Cell Biol 196: 363–374.
Oertner TG, Matus A. 2005. Calcium regulation of actin dynamics in dendritic spines. Cell Calcium
37: 477–482.
Padrick SB, Rosen MK. 2010. Physical mechanisms of signal integration by WASP family proteins.
Annu Rev Biochem 79: 707–735.
Palecek SP, Loftus JC, Ginsberg MH, Lauffenburger DA, Horwitz AF. 1997. Integrin-ligand binding
properties govern cell migration speed through cell-substratum adhesiveness. Nature 385: 537–540.
Parent CA, Devreotes PN. 1999. A cell’s sense of direction. Science 284: 765–770.
Parent CA, Blacklock BJ, Froehlich WM, Murphy DB, Devreotes PN. 1998. G protein signaling events
are activated at the leading edge of chemotactic cells. Cell 95: 81–91.
Parsons JT. 2003. Focal adhesion kinase: The first ten years. J Cell Sci 116: 1409–1416.
Parsons JT, Horwitz AR, Schwartz MA. 2010. Cell adhesion: Integrating cytoskeletal dynamics and
cellular tension. Nat Rev Mol Cell Biol 11: 633–643.
Paul AS, Pollard TD. 2009. Review of the mechanism of processive actin filament elongation by
formins. Cell Motil Cytoskeleton 66: 606–617.
Petrie RJ, Gavara N, Chadwick RS, Yamada KM. 2012. Nonpolarized signaling reveals two distinct
modes of 3D cell migration. J Cell Biol 197: 439–455.
Pollard TD, Borisy GG. 2003. Cellular motility driven by assembly and disassembly of actin filaments.
Cell 112: 453–465.
Pollitt AY, Insall RH. 2009. WASP and SCAR/WAVE proteins: The drivers of actin assembly. J Cell
Sci 122: 2575–2578.
Ponti A, Machacek M, Gupton SL, Waterman-Storer CM, Danuser G. 2004. Two distinct actin
networks drive the protrusion of migrating cells. Science 305: 1782–1786.
Provenzano PP, Eliceiri KW, Keely PJ. 2009. Shining new light on 3D cell motility and the metastatic
process. Trends Cell Biol 19: 638–648.
Ridley AJ. 2006. Rho GTPases and actin dynamics in membrane protrusions and vesicle trafficking.
Trends Cell Biol 16: 522–529.
Ridley AJ. 2011. Life at the leading edge. Cell 145: 1012–1022.
Ridley AJ, Schwartz MA, Burridge K, Firtel RA, Ginsberg MH, Borisy G, Parsons JT, Horwitz AR.
2003. Cell migration: Integrating signals from front to back. Science 302: 1704–1709.
Robinson DN, Spudich JA. 2004. Mechanics and regulation of cytokinesis. Curr Opin Cell Biol 16:
182–188.
Ruusala A, Aspenstrom P. 2008. The atypical Rho GTPase Wrch1 collaborates with the nonreceptor
tyrosine kinases Pyk2 and Src in regulating cytoskeletal dynamics. Mol Cell Biol 28: 1802–1814.
Sasaki AT, Firtel RA. 2009. Spatiotemporal regulation of Ras-GTPases during chemotaxis. Methods
Mol Biol 571: 333–348.
Sawada Y, Tamada M, Dubin-Thaler BJ, Cherniavskaya O, Sakai R, Tanaka S, Sheetz MP. 2006.
Force sensing by mechanical extension of the Src family kinase substrate p130Cas. Cell 127: 1015–
1026.
Schmidt S, Friedl P. 2010. Interstitial cell migration: Integrin-dependent and alternative adhesion
mechanisms. Cell Tissue Res 339: 83–92.
Schneider IC, Haugh JM. 2006. Mechanisms of gradient sensing and chemotaxis: Conserved pathways,
diverse regulation. Cell Cycle 5: 1130–1134.
Schwartz MA. 2010. Integrins and extracellular matrix in mechanotransduction. Cold Spring Harb
Perspect Biol 2: a005066.
Sepulveda JL, Gkretsi V, Wu C. 2005. Assembly and signaling of adhesion complexes. Curr Top Dev
Biol 68: 183–225.
* Sever R, Brugge JS. 2014. Signal transduction in cancer. Cold Spring Harb Perspect Med doi:
10.1101/cshperspect.a006098.
Shattil SJ, Kim C, Ginsberg MH. 2010. The final steps of integrin activation: The end game. Nat Rev
Mol Cell Biol 11: 288–300.
Shewan A, Eastburn DJ, Mostov K. 2011. Phosphoinositides in cell architecture. Cold Spring Harb
Perspect Biol doi: 10.1101/cshperspect.a004796.
Slessareva JE, Dohlman HG. 2006. G protein signaling in yeast: New components, new connections,
new compartments. Science 314: 1412–1413.
Small JV, Resch GP. 2005. The comings and goings of actin: Coupling protrusion and retraction in cell
motility. Curr Opin Cell Biol 17: 517–523.
Small JV, Stradal T, Vignal E, Rottner K. 2002. The lamellipodium: Where motility begins. Trends
Cell Biol 12: 112–120.
Swaney KF, Huang CH, Devreotes PN. 2010. Eukaryotic chemotaxis: A network of signaling pathways
controls motility, directional sensing, and polarity. Annu Rev Biophys 278: 20445–20448.
Takeuchi H, Higashiyama T. 2011. Attraction of tip-growing pollen tubes by the female gametophyte.
Curr Opin Plant Biol 14: 614–621.
Tada T, Sheng M. 2006. Molecular mechanisms of dendritic spine morphogenesis. Curr Opin
Neurobiol 16: 95–101.
Tang M, Iijima M, Kamimura Y, Chen L, Long Y, Devreotes PN. 2011a. Disruption of PKB signaling
restores polarity to cells lacking tumor suppressor PTEN. Mol Biol Cell 22: 437–447.
Tang M, Iijima M, Devreotes P. 2011b. Generation of cells that ignore the effects of PIP3 on
cytoskeleton. Cell Cycle 10: 2817–2818.
Tomar A, Schlaepfer DD. 2009. Focal adhesion kinase: Switching between GAPs and GEFs in the
regulation of cell motility. Curr Opin Cell Biol 21: 676–683.
Trinkaus JP. 1969. Cells into organs: The forces that shape the embryo, p. 215. Prentice-Hall,
Englewood Cliffs, NJ.
Turner CE, West KA, Brown MC. 2001. Paxillin-ARF GAP signaling and the cytoskeleton. Curr Opin
Cell Biol 13: 593–599.
Tybulewicz VL, Ardouin L, Prisco A, Reynolds LF. 2003. Vav1: A key signal transducer downstream
of the TCR. Immunol Rev 192: 42–52.
van Galen EJ, Ramakers GJ. 2005. Rho proteins, mental retardation and the neurobiological basis of
intelligence. Prog Brain Res 147: 295–317.
Van Haastert PJ, Devreotes PN. 2004. Chemotaxis: Signalling the way forward. Nat Rev Mol Cell Biol
5: 626–634.
Van Keymeulen A, Wong K, Knight ZA, Govaerts C, Hahn KM, Shokat KM, Bourne HR. 2006. To
stabilize neutrophil polarity, PIP3 and Cdc42 augment RhoA activity at the back as well as signals
at the front. J Cell Biol 174: 437–445.
Veltman DM, Keizer-Gunnik I, Van Haastert PJ. 2008. Four key signaling pathways mediating
chemotaxis in Dictyostelium discoideum. J Cell Biol 180: 747–753.
Vicente-Manzanares M, Sancho D, Yáñez-Mó M, Sánchez-Madrid F. 2002. The leukocyte
cytoskeleton in cell migration and immune interactions. Int Rev Cytol 216: 233–289.
Vicente-Manzanares M, Koach MA, Whitmore L, Lamers ML, Horwitz AF. 2008. Segregation and
activation of myosin IIB creates a rear in migrating cells. J Cell Biol 183: 543–554.
Vicente-Manzanares M, Ma X, Adelstein RS, Horwitz AR. 2009. Non-muscle myosin II takes centre
stage in cell adhesion and migration. Nat Rev Mol Cell Biol 10: 778–790.
Vicente-Manzanares M, Newell-Litwa K, Bachir AI, Whitmore LA, Horwitz AR. 2011. Myosin
IIA/IIB restrict adhesive and protrusive signaling to generate front-back polarity in migrating cells.
J Cell Biol 193: 381–396.
Wang Y-L. 2007. Flux at focal adhesions: Slippage clutch, mechanical gauge, or signal depot. Sci
STKE 2007: e10.
Wang F, Herzmark P, Weiner OD, Srinivasan S, Servant G, Bourne HR. 2002. Lipid products of
PI(3)Ks maintain persistent cell polarity and directed motility in neutrophils. Nature Cell Biology 4:
513–518.
Webb DJ, Donais K, Whitmore LA, Thomas SM, Turner CE, Parsons JT, Horwitz AF. 2004. FAK-Src
signalling through paxillin, ERK and MLCK regulates adhesion disassembly. Nat Cell Biol 6: 154–
161.
Weiner OD, Neilsen PO, Prestwich GD, Kirschner MW, Cantley LC, Bourne HR. 2002. A PtdInsP(3)-
and Rho GTPase-mediated positive feedback loop regulates neutrophil polarity. Nat Cell Biol 4:
509–513.
Weiner OD, Marganski WA, Wu LF, Altschuler SJ, Kirschner MW. 2007. An actin-based wave
generator organizes cell motility. PLoS Biol 5: e221.
Welch HC, Coadwell WJ, Ellson CD, Ferguson GJ, Andrews SR, Erdjument-Bromage H, Tempst P,
Hawkins PT, Stephens LR. 2002. P-Rex1, a PtdIns(3,4,5)P3- and Gβγ-regulated guanine-nucleotide
exchange factor for Rac. Cell 108: 809–821.
Xiong Y, Huang C-H, Iglesias PA, Devreotes PN. 2010. Cells navigate with a local-excitation, global-
inhibition-biased excitable network. Proc Natl Acad Sci 107: 17079–17086.
Yamaguchi H, Condeelis J. 2007. Regulation of the actin cytoskeleton in cancer cell migration and
invasion. Biochim Biophys Acta 1773: 642–652.
Yoo SK, Deng Q, Cavnar PJ, Wu YI, Hahn KM, Huttenlocher A. 2010. Differential regulation of
protrusion and polarity by PI3K during neutrophil motility in live zebrafish. Dev Cell 18: 226–236.
Zaidel-Bar R, Geiger B. 2010. The switchable integrin adhesome. J Cell Sci 123: 1385–1388.
Zaidel-Bar R, Itzkovitz S, Ma’ayan A, Iyengar R, Geiger B. 2007a. Functional atlas of the integrin
adhesome. Nat Cell Biol 9: 858–867.
Zaidel-Bar R, Milo R, Kam Z, Geiger B. 2007b. A paxillin tyrosine phosphorylation switch regulates
the assembly and form of cell-matrix adhesions. J Cell Sci 120: 137–148.
Zhang H, Webb D, Asmussen J, Niu H, Horwitz AF. 2005. A GIT1/PIX/Rac/PAK signaling module
regulates spine morphogenesis and synapse formation through MLC. J Neurosci 25: 3379–3388.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a005959
CHAPTER 9
SUMMARY
Outline
1 Introduction
2 The polarization machinery
3 Factors that control polarity protein localization
4 Polarity signaling through PAR3–PAR6–APKC
5 Conclusion
References
1 INTRODUCTION
The asymmetric distribution of proteins, lipids, and RNAs is necessary for
cell fate determination, differentiation, and a multitude of specialized cell
functions that underlie morphogenesis (St Johnston 2005; Gonczy 2008;
Knoblich 2008; Macara and Mili 2008; Martin-Belmonte and Mostov 2008).
The establishment of cell polarity can be dissected into three primary
processes: (1) breaking symmetry, either through extrinsic cues or
stochastically; (2) establishing spatial organization through signal
transduction; and (3) amplifying and maintaining the polarized state through
feedback loops (Fig. 1). Even single-celled organisms such as budding yeast
are polarized and engage sophisticated signaling mechanisms to initiate and
organize asymmetric cell divisions. Higher organisms use polarity to build
diverse cell types, such as neurons and epithelial cells in animals or stomatal
cells in plants. Polarity spatially segregates important cellular functions from
one another—for instance, in neurons, it separates synaptic inputs (along
dendrites) from signaling outputs (along the axons). Epithelial cell polarity
separates the apical membrane, which is specialized for interactions with the
external environment, from the baso-lateral membrane, which contacts
extracellular matrix or other cell types. In some epithelia a barrier called the
tight junction separates the two membrane regions and prevents the
intercellular diffusion of material across the epithelial sheet. Once
established, cell polarity is often stable for the lifetime of the cell, as in
neurons, but it can also be dynamic, for example, during development, when
neural crest cells lose their epithelial character and become mesenchymal
(this is termed the epithelial mesenchymal transition, EMT).
Figure 1. (A) Extrinsic signals normally are responsible for driving cell polarization, although it can
also occur spontaneously under certain conditions. (B) Cell polarization is established by signal
transduction pathways that spatially segregate different regions of the cell, especially the cell cortex,
and this organization is reinforced and maintained by positive-feedback loops.
Figure 2. (A) Schematic showing domain structures of Par polarity proteins and their interactions. Phox
and Bem1 domain (PB1), forms homodimers and heterodimers; zinc finger domain (Zn); PSD95, Dlg1,
ZO-1 domain (PDZ), binds other PDZ domains and carboxy-terminal peptide motifs; conserved region
domain (CR1), forms homo-oligomers; atypical protein kinase C binding domain (aPKCBD);
ubiquitin-binding-associated domain (UBA); kinase-associated domain (KA). (B) The different
distributions of these polarity proteins in an epithelial cell and a neuroblast stem cell, together with the
localization of other interacting proteins. Note that whereas in neuroblasts all of the polarity proteins
form a complex (the “Par complex”) at the apical cortex, this is not the case in epithelial cells, in which
Par3 is not associated with Par6 and aPKC but is associated instead with the tight junction complex.
The orientation of the mitotic spindle is controlled by the Par proteins and is different in neuroblasts
(vertical) versus epithelial cells (horizontal). This difference reflects the distinct functions of polarity in
the two cell types: segregation of cell fate determinants into only one daughter cell in the neuroblast
versus formation of a polarized sheet of cells by the epithelium.
3.2 Oligomerization
The amino-terminal conserved region 1 (CR1) is necessary for self-
association of Par3 into higher-order complexes (Fig. 2) (Benton and
Johnston 2003a; Mizuno et al. 2003; Feng et al. 2007). It is essential but not
sufficient for membrane attachment. How the oligomerization of Par3
maintains the protein at the plasma membrane is unclear, but oligomerization
might complement weak phosphoinositide binding by increasing avidity (Fig.
4).
Figure 4. Mechanisms for the transport, cortical association, and anchoring of Par3 in mammalian
cells. It is not yet known if all of these mechanisms operate in any one cell type, and additional
processes, such as RNA localization, might play roles in certain circumstances. Junctional adhesion
molecule (JAM) is shown as an example of a transmembrane protein to which Par3 can be anchored,
but others exist, such as the neurotrophin receptor, p75NTR, in mammalian Schwann cells (Chan et al.
2006). PP1α is a phosphatase.
5 CONCLUSION
Understanding cell polarization is one of the major goals of cell biology and
will inevitably have a broad impact on research into diseases such as cancer
and neurological degeneration. A complicated web of signaling systems
surrounds and intersects with the polarity machinery, yet we still understand
very little about what the Par proteins do, how they are localized, how their
various interactions are regulated, and which signaling components operate in
which contexts. After all, the organization of a polarized cell is a formidably
complicated process that involves cytoskeletal remodeling, membrane traffic,
RNA localization, and protein complex assembly and disassembly, with
feedback to gene expression and protein turnover. It is conceivable that the
Par proteins participate in all of these processes, either directly or indirectly.
It is worth remembering, however, that although the polarity machinery can
work in a cell-autonomous fashion, the Par genes do not exist in any
unicellular organism, which suggests that a key role for the Par proteins is to
facilitate, mediate, or interpret cell–cell and cell–matrix interactions. The
tissue context might, therefore, be expected to modulate Par protein
functions. It will be interesting to determine whether regulation of these
interactions controls morphogenesis and to what extent differential Par
function contributes to phenotypic variation both between different cell types
in one organism and between species.
REFERENCES
*Reference is in this book.
Aceto D, Beers M, Kemphues KJ. 2006. Interaction of PAR-6 with CDC-42 is required for
maintenance but not establishment of PAR asymmetry in C. elegans. Dev Biol 299: 386–397.
Atwood SX. Prehoda KE. 2009. aPKC phosphorylates Miranda to polarize fate determinants during
neuroblast asymmetric cell division. Curr Biol 19: 723–729.
Atwood SX, Chabu C, Penkert RR, Doe CQ, Prehoda KE. 2007. Cdc42 acts downstream of Bazooka to
regulate neuroblast polarity through Par-6 aPKC. J Cell Sci 120: 3200–3206.
Baas AF, Kuipers J, van der Wel NN, Batlle E, Koerten HK, Peters PJ, Clevers HC. 2004. Complete
polarization of single intestinal epithelial cells upon activation of LKB1 by STRAD. Cell 116: 457–
466.
Beatty A, Morton D, Kemphues K. 2010. The C. elegans homolog of Drosophila Lethal giant larvae
functions redundantly with PAR-2 to maintain polarity in the early embryo. Development 137:
3995–4004.
Benton R, Johnston DS. 2003a. A conserved oligomerization domain in Drosophila Bazooka/PAR-3 is
important for apical localization and epithelial polarity. Curr Biol 13: 1330–1334.
Benton R, Johnston DS. 2003b. Drosophila PAR-1 and 14-3-3 inhibit Bazooka/PAR-3 to establish
complementary cortical domains in polarized cells. Cell 115: 691–704.
Betschinger J, Mechtler K, Knoblich JA. 2003. The Par complex directs asymmetric cell division by
phosphorylating the cytoskeletal protein Lgl. Nature 422: 326–330.
Brabletz S, Brabletz T. 2010. The ZEB/miR-200 feedback loop—a motor of cellular plasticity in
development and cancer? EMBO Rep 11: 670–677.
Bruewer M, Hopkins AM, Hobert ME, Nusrat A, Madara JL. 2004. RhoA, Rac1, and Cdc42 exert
distinct effects on epithelial barrier via selective structural and biochemical modulation of
junctional proteins and F-actin. Am J Physiol Cell Physiol 287: C327–C335.
Büther K, Plaas C, Barnekow A, Kremerskothen J. 2004. KIBRA is a novel substrate for protein kinase
Cζ. Biochem Biophys Res Commun 317: 703–707.
Butty AC, Perrinjaquet N, Petit A, Jaquenoud M, Segall JE, Hofmann K, Zwahlen C, Peter M. 2002. A
positive feedback loop stabilizes the guanine-nucleotide exchange factor Cdc24 at sites of
polarization. EMBO J 21: 1565–1576.
Chan JR, Jolicoeur C, Yamauchi J, Elliott J, Fawcett JP, Ng BK, Cayouette M. 2006. The polarity
protein Par-3 directly interacts with p75NTR to regulate myelination. Science 314: 832–836.
Charest PG, Firtel RA. 2006. Feedback signaling controls leading-edge formation during chemotaxis.
Curr Opin Genet Dev 16: 339–347.
Chen X, Macara IG. 2005. Par-3 controls tight junction assembly through the Rac exchange factor
Tiam1. Nat Cell Biol 7: 262–269.
Chen CL, Gajewski KM, Hamaratoglu F, Bossuyt W, Sansores-Garcia L, Tao C, Halder G. 2010. The
apical–basal cell polarity determinant Crumbs regulates Hippo signaling in Drosophila. Proc Natl
Acad Sci 107: 15810–15815.
Courbard JR, Djiane A, Wu J, Mlodzik M. 2009. The apical/basal-polarity determinant Scribble
cooperates with the PCP core factor Stbm/Vang and functions as one of its effectors. Dev Biol 333:
67–77.
* Devreotes P, Horwitz AF. 2012. Cell migration and chemotaxis. Cold Spring Harb Perspect Biol 4:
a005959.
Djiane A, Yogev S, Mlodzik M. 2005. The apical determinants aPKC and dPatj regulate Frizzled-
dependent planar cell polarity in the Drosophila eye. Cell 121: 621–631.
Dupont S, Morsut L, Aragona M, Enzo E, Giulitti S, Cordenonsi M, Zanconato F, Le Digabel J,
Forcato M, Bicciato S, et al. 2011. Role of YAP/TAZ in mechanotransduction. Nature 474: 179–
183.
Ebnet K, Schulz CU, Meyer Zu Brickwedde MK, Pendl GG, Vestweber D. 2000. Junctional adhesion
molecule interacts with the PDZ domain-containing proteins AF-6 and ZO-1. J Biol Chem 275:
27979–27988.
Etemad-Moghadam B, Guo S, Kemphues KJ. 1995. Asymmetrically distributed PAR-3 protein
contributes to cell polarity and spindle alignment in early C. elegans embryos. Cell 83: 743–752.
Feng W, Wu H, Chan LN, Zhang M. 2007. The Par-3 NTD adopts a PB1-like structure required for
Par-3 oligomerization and membrane localization. EMBO J 26: 2786–2796.
Gao L, Joberty G, Macara IG. 2002. Assembly of epithelial tight junctions is negatively regulated by
Par6. Curr Biol 12: 221–225.
Garrard SM, Capaldo CT, Gao L, Rosen MK, Macara IG, Tomchick DR. 2003. Structure of Cdc42 in a
complex with the GTPase-binding domain of the cell polarity protein, Par6. EMBO J 22: 1125–
1133.
Georgiou M, Baum B. 2010. Polarity proteins and Rho GTPases cooperate to spatially organise
epithelial actin-based protrusions. J Cell Sci 123: 1089–1098.
Gladden AB, Hebert AM, Schneeberger EE, McClatchey AI. 2010. The NF2 tumor suppressor, Merlin,
regulates epidermal development through the establishment of a junctional polarity complex. Dev
Cell 19: 727–739.
Goehring NW, Hoege C, Grill SW, Hyman AA. 2011. PAR proteins diffuse freely across the anterior–
posterior boundary in polarized C. elegans embryos. J Cell Biol 193: 583–594.
Goldstein B, Macara IG. 2007. The PAR proteins: Fundamental players in animal cell polarization. Dev
Cell 13: 609–622.
Gonczy P. 2008. Mechanisms of asymmetric cell division: Flies and worms pave the way. Nat Rev Mol
Cell Biol 9: 355–366.
Grusche FA, Richardson HE, Harvey KF. 2010. Upstream regulation of the Hippo size control
pathway. Curr Biol 20: R574–R582.
Grzeschik NA, Parsons LM, Allott ML, Harvey KF, Richardson HE. 2010. Lgl, aPKC, and Crumbs
regulate the Salvador/Warts/Hippo pathway through two distinct mechanisms. Curr Biol 20: 573–
581.
Guo S, Kemphues KJ. 1995. par-1, a gene required for establishing polarity in C. elegans embryos,
encodes a putative Ser/Thr kinase that is asymmetrically distributed. Cell 81: 611–620.
Hao Y, Boyd L, Seydoux G. 2006. Stabilization of cell polarity by the C. elegans RING protein PAR-2.
Dev Cell 10: 199–208.
Harris TJ, Peifer M. 2005. The positioning and segregation of apical cues during epithelial polarity
establishment in Drosophila. J Cell Biol 170: 813–823.
* Harvey KF, Hariharan IK. 2012. The Hippo pathway. Cold Spring Harb Perspect Biol 4: a011288.
Hengst U, Deglincerti A, Kim HJ, Jeon NL, Jaffrey SR. 2009. Axonal elongation triggered by stimulus-
induced local translation of a polarity complex protein. Nat Cell Biol 11: 1024–1030.
Hirano Y, Yoshinaga S, Takeya R, Suzuki NN, Horiuchi M, Kohjima M, Sumimoto H, Inagaki F.
2005. Structure of a cell polarity regulator, a complex between atypical PKC and Par6 PB1
domains. J Biol Chem 280: 9653–9661.
Hoege C, Constantinescu AT, Schwager A, Goehring NW, Kumar P, Hyman AA. 2010. LGL can
partition the cortex of one-cell Caenorhabditis elegans embryos into two domains. Curr Biol 20:
1296–1303.
Horne-Badovinac S, Bilder D. 2008. Dynein regulates epithelial polarity and the apical localization of
stardust A mRNA. PLoS Genet 4: e8.
Hung TJ, Kemphues KJ. 1999. PAR-6 is a conserved PDZ domain-containing protein that colocalizes
with PAR-3 in Caenorhabditis elegans embryos. Development 126: 127–135.
Hurd TW, Fan S, Liu C-J, Kweon HK, Hakansson K, Margolis B. 2003a. Phosphorylation-dependent
binding of 14-3-3 to the polarity protein Par3 regulates cell polarity in mammalian epithelia. Curr
Biol 13: 2082–2090.
Hurd TW, Gao L, Roh MH, Macara IG, Margolis B. 2003b. Direct interaction of two polarity
complexes implicated in epithelial tight junction assembly. Nat Cell Biol 5: 137–142.
Hurov JB, Watkins JL, Piwnica-Worms H. 2004. Atypical PKC phosphorylates PAR-1 kinases to
regulate localization and activity. Curr Biol 14: 736–741.
Ishiuchi T, Takeichi M. 2011. Willin and Par3 cooperatively regulate epithelial apical constriction
through aPKC-mediated ROCK phosphorylation. Nat Cell Biol 13: 860–866.
Itoh M, Sasaki H, Furuse M, Ozaki H, Kita T, Tsukita S. 2001. Junctional adhesion molecule (JAM)
binds to PAR-3: A possible mechanism for the recruitment of PAR-3 to tight junctions. J Cell Biol
154: 491–497.
Izumi Y, Hirose T, Tamai Y, Hirai S, Nagashima Y, Fujimoto T, Tabuse Y, Kemphues KJ, Ohno S.
1998. An atypical PKC directly associates and colocalizes at the epithelial tight junction with ASIP,
a mammalian homologue of Caenorhabditis elegans polarity protein PAR-3. J Cell Biol 143: 95–
106.
Joberty G, Petersen C, Gao L, Macara IG. 2000. The cell-polarity protein Par6 links Par3 and atypical
protein kinase C to Cdc42. Nat Cell Biol 2: 531–539.
Kemphues KJ, Priess JR, Morton DG, Cheng NS. 1988. Identification of genes required for
cytoplasmic localization in early C. elegans embryos. Cell 52: 311–320.
Knoblich JA. 2008. Mechanisms of asymmetric stem cell division. Cell 132: 583–597.
Krahn MP, Buckers J, Kastrup L, Wodarz A. 2010. Formation of a Bazooka–Stardust complex is
essential for plasma membrane polarity in epithelia. J Cell Biol 190: 751–760.
Lee HS, Nishanian TG, Mood K, Bong YS, Daar IO. 2008. EphrinB1 controls cell–cell junctions
through the Par polarity complex. Nat Cell Biol 10: 979–986.
Levitan DJ, Boyd L, Mello CC, Kemphues KJ, Stinchcomb DT. 1994. par-2, a gene required for
blastomere asymmetry in Caenorhabditis elegans, encodes zinc-finger and ATP-binding motifs.
Proc Natl Acad Sci 91: 6108–6112.
Li Z, Wang L, Hays TS, Cai Y. 2008. Dynein-mediated apical localization of crumbs transcripts is
required for Crumbs activity in epithelial polarity. J Cell Biol 180: 31–38.
Lin D, Edwards AS, Fawcett JP, Mbamalu G, Scott JD, Pawson T. 2000. A mammalian PAR-3–PAR-6
complex implicated in Cdc42/Rac1 and aPKC signalling and cell polarity. Nat Cell Biol 2: 540–
547.
Ling C, Zheng Y, Yin F, Yu J, Huang J, Hong Y, Wu S, Pan D. 2010. The apical transmembrane
protein Crumbs functions as a tumor suppressor that regulates Hippo signaling by binding to
Expanded. Proc Natl Acad Sci 107: 10532–10537.
Lu H, Bilder D. 2005. Endocytic control of epithelial polarity and proliferation in Drosophila. Nat Cell
Biol 7: 1232–1239.
Macara IG, Mili S. 2008. Polarity and differential inheritance—universal attributes of life? Cell 135:
801–812.
Martin-Belmonte F, Mostov K. 2008. Regulation of cell polarity during epithelial morphogenesis. Curr
Opin Cell Biol 20: 227–234.
Martin-Belmonte F, Gassama A, Datta A, Yu W, Rescher U, Gerke V, Mostov K. 2007. PTEN-
mediated apical segregation of phosphoinositides controls epithelial morphogenesis through Cdc42.
Cell 128: 383–397.
Mayer B, Emery G, Berdnik D, Wirtz-Peitz F, Knoblich JA. 2005. Quantitative analysis of protein
dynamics during asymmetric cell division. Curr Biol 15: 1847–1854.
Mertens AE, Rygiel TP, Olivo C, van der Kammen R, Collard JG. 2005. The Rac activator Tiam1
controls tight junction biogenesis in keratinocytes through binding to and activation of the Par
polarity complex. J Cell Biol 170: 1029–1037.
Mizuno K, Suzuki A, Hirose T, Kitamura K, Kutsuzawa K, Futaki M, Amano Y, Ohno S. 2003. Self-
association of PAR-3-mediated by the conserved N-terminal domain contributes to the development
of epithelial tight junctions. J Biol Chem 278: 31240–31250.
Morais-de-Sa E, Mirouse V, St Johnston D. 2010. aPKC phosphorylation of Bazooka defines the
apical/lateral border in Drosophila epithelial cells. Cell 141: 509–523.
Morton DG, Shakes DC, Nugent S, Dichoso D, Wang W, Golden A, Kemphues KJ. 2002. The
Caenorhabditis elegans par-5 gene encodes a 14-3-3 protein required for cellular asymmetry in the
early embryo. Dev Biol 241: 47–58.
Motegi F, Sugimoto A. 2006. Sequential functioning of the ECT-2 Rho-GEF, RHO-1 and CDC-42
establishes cell polarity in Caenorhabditis elegans embryos. Nat Cell Biol 8: 978–985.
Munro EM. 2006. PAR proteins and the cytoskeleton: A marriage of equals. Curr Opin Cell Biol 18:
86–94.
Nagai-Tamai Y, Mizuno K, Hirose T, Suzuki A, Ohno S. 2002. Regulated protein–protein interaction
between aPKC and PAR-3 plays an essential role in the polarization of epithelial cells. Genes Cells
7: 1161–1171.
Nakayama M, Goto TM, Sugimoto M, Nishimura T, Shinagawa T, Ohno S, Amano M, Kaibuchi K.
2008. Rho-kinase phosphorylates PAR-3 and disrupts PAR complex formation. Dev Cell 14: 205–
215.
Nishimura T, Kaibuchi K. 2007. Numb controls integrin endocytosis for directional cell migration with
aPKC and PAR-3. Dev Cell 13: 15–28.
Nishimura T, Yamaguchi T, Kato K, Yoshizawa M, Nabeshima YI, Ohno S, Hoshino M, Kaibuchi K.
2005. PAR-6–PAR-3 mediates Cdc42-induced Rac activation through the Rac GEFs STEF/Tiam1.
Nat Cell Biol 7: 270–277.
Nobes CD, Hall A. 1999. Rho GTPases control polarity, protrusion, and adhesion during cell
movement. J Cell Biol 144: 1235–1244.
Ossipova O, Dhawan S, Sokol S, Green JB. 2005. Distinct PAR-1 proteins function in different
branches of Wnt signaling during vertebrate development. Dev Cell 8: 829–841.
Ozdamar B, Bose R, Barrios-Rodiles M, Wang HR, Zhang Y, Wrana JL. 2005. Regulation of the
polarity protein Par6 by TGFβ receptors controls epithelial cell plasticity. Science 307: 1603–1609.
Pegtel DM, Ellenbroek SI, Mertens AE, van der Kammen RA, de Rooij J, Collard JG. 2007. The par–
tiam1 complex controls persistent migration by stabilizing microtubule-dependent front–rear
polarity. Curr Biol 17: 1623–1634.
Robinson BS, Huang J, Hong Y, Moberg KH. 2010. Crumbs regulates Salvador/Warts/Hippo signaling
in Drosophila via the FERM-domain protein Expanded. Curr Biol 20: 582–590.
Rorth P. 2009. Collective cell migration. Annu Rev Cell Dev Biol 25: 407–429.
Rosse C, Formstecher E, Boeckeler K, Zhao Y, Kremerskothen J, White MD, Camonis JH, Parker PJ.
2009. An aPKC–exocyst complex controls paxillin phosphorylation and migration through
localised JNK1 activation. PLoS Biol 7: e1000235.
Schlegelmilch K, Mohseni M, Kirak O, Pruszak J, Rodriguez JR, Zhou D, Kreger BT, Vasioukhin V,
Avruch J, Brummelkamp TR, et al. 2011. Yap1 acts downstream of α-catenin to control epidermal
proliferation. Cell 144: 782–795.
Schlessinger K, McManus EJ, Hall A. 2007. Cdc42 and noncanonical Wnt signal transduction
pathways cooperate to promote cell polarity. J Cell Biol 178: 355–361.
Schneider SQ, Bowerman B. 2003. Cell polarity and the cytoskeleton in the Caenorhabditis elegans
zygote. Annu Rev Genet 37: 221–249.
* Sever R, Brugge JS. 2014. Signal transduction in cancer. Cold Spring Harb Perspect Med doi:
10.1101/cshperspect.a006098.
Silvis MR, Kreger BT, Lien WH, Klezovitch O, Rudakova GM, Camargo FD, Lantz DM, Seykora JT,
Vasioukhin V. 2011. α-Catenin is a tumor suppressor that controls cell accumulation by regulating
the localization and activity of the transcriptional coactivator Yap1. Sci Signal 4: ra33.
Simons M, Mlodzik M. 2008. Planar cell polarity signaling: From fly development to human disease.
Annu Rev Genet 42: 517–540.
Slaughter BD, Smith SE, Li R. 2009. Symmetry breaking in the life cycle of the budding yeast. Cold
Spring Harb Perspect Biol 1: a003384.
Smith CA, Lau KM, Rahmani Z, Dho SE, Brothers G, She YM, Berry DM, Bonneil E, Thibault P,
Schweisguth F, et al. 2007. aPKC-mediated phosphorylation regulates asymmetric membrane
localization of the cell fate determinant Numb. Embo J 26: 468–480.
St Johnston D. 2005. Moving messages: The intracellular localization of mRNAs. Nat Rev Mol Cell
Biol 6: 363–375.
Sun TQ, Lu B, Feng JJ, Reinhard C, Jan YN, Fantl WJ, Williams LT. 2001. PAR-1 is a Dishevelled-
associated kinase and a positive regulator of Wnt signalling. Nat Cell Biol 3: 628–636.
Tabuse Y, Izumi Y, Piano F, Kemphues KJ, Miwa J, Ohno S. 1998. Atypical protein kinase C
cooperates with PAR-3 to establish embryonic polarity in Caenorhabditis elegans. Development
125: 3607–3614.
Takekuni K, Ikeda W, Fujito T, Morimoto K, Takeuchi M, Monden M, Takai Y. 2003. Direct binding
of cell polarity protein PAR-3 to cell–cell adhesion molecule nectin at neuroepithelial cells of
developing mouse. J Biol Chem 278: 5497–5500.
Tong Z, Gao XD, Howell AS, Bose I, Lew DJ, Bi E. 2007. Adjacent positioning of cellular structures
enabled by a Cdc42 GTPase-activating protein-mediated zone of inhibition. J Cell Biol 179: 1375–
1384.
Tsuji T, Ohta Y, Kanno Y, Hirose K, Ohashi K, Mizuno K. 2010. Involvement of p114–RhoGEF and
Lfc in Wnt-3a- and Dishevelled-induced RhoA activation and neurite retraction in N1E-115 mouse
neuroblastoma cells. Mol Biol Cell 21: 3590–3600.
Varelas X, Samavarchi-Tehrani P, Narimatsu M, Weiss A, Cockburn K, Larsen BG, Rossant J, Wrana
JL. 2010. The Crumbs complex couples cell density sensing to Hippo-dependent control of the
TGF-β–SMAD pathway. Dev Cell 19: 831–844.
Vidal M, Salavaggione L, Ylagan L, Wilkins M, Watson M, Weilbaecher K, Cagan R. 2010. A role for
the epithelial microenvironment at tumor boundaries: Evidence from Drosophila and human
squamous cell carcinomas. Am J Pathol 176: 3007–3014.
Wang HR, Zhang Y, Ozdamar B, Ogunjimi AA, Alexandrova E, Thomsen GH, Wrana JL. 2003.
Regulation of cell polarity and protrusion formation by targeting RhoA for degradation. Science
302: 1775–1779.
Watkins JL, Lewandowski KT, Meek SE, Storz P, Toker A, Piwnica-Worms H. 2008. Phosphorylation
of the Par-1 polarity kinase by protein kinase D regulates 14-3-3 binding and membrane
association. Proc Natl Acad Sci 105: 18378–18383.
Watts JL, Morton DG, Bestman J, Kemphues KJ. 2000. The C. elegans par-4 gene encodes a putative
serine-threonine kinase required for establishing embryonic asymmetry. Development 127: 1467–
1475.
Wei SY, Escudero LM, Yu F, Chang LH, Chen LY, Ho YH, Lin CM, Chou CS, Chia W, Modolell J, et
al. 2005. Echinoid is a component of adherens junctions that cooperates with DE-Cadherin to
mediate cell adhesion. Dev Cell 8: 493–504.
Wirtz-Peitz F, Nishimura T, Knoblich JA. 2008. Linking cell cycle to asymmetric division: Aurora-A
phosphorylates the Par complex to regulate Numb localization. Cell 135: 161–173.
Wodarz A, Ramrath A, Grimm A, Knust E. 2000. Drosophila atypical protein kinase C associates with
Bazooka and controls polarity of epithelia and neuroblasts. J Cell Biol 150: 1361–1374.
Wu H, Feng W, Chen J, Chan LN, Huang S, Zhang M. 2007. PDZ domains of Par-3 as potential
phosphoinositide signaling integrators. Mol Cell 28: 886–898.
Wu M, Pastor-Pareja JC, Xu T. 2010. Interaction between Ras(V12) and scribbled clones induces
tumour growth and invasion. Nature 463: 545–548.
Xiong Y, Huang CH, Iglesias PA, Devreotes PN. 2010. Cells navigate with a local-excitation, global-
inhibition-biased excitable network. Proc Natl Acad Sci 107: 17079–17086.
Yamanaka T, Horikoshi Y, Suzuki A, Sugiyama Y, Kitamura K, Maniwa R, Nagai Y, Yamashita A,
Hirose T, Ishikawa H, et al. 2001. PAR-6 regulates aPKC activity in a novel way and mediates
cell–cell contact-induced formation of the epithelial junctional complex. Genes Cells 6: 721–731.
Yoshihama Y, Sasaki K, Horikoshi Y, Suzuki A, Ohtsuka T, Hakuno F, Takahashi S, Ohno S, Chida K.
2010. KIBRA suppresses apical exocytosis through inhibition of aPKC kinase activity in epithelial
cells. Curr Biol 21: 705–711.
Zeitler J, Hsu CP, Dionne H, Bilder D. 2004. Domains controlling cell polarity and proliferation in the
Drosophila tumor suppressor Scribble. J Cell Biol 167: 1137–1146.
Zhang H, Macara IG. 2006. The polarity protein PAR-3 and TIAM1 cooperate in dendritic spine
morphogenesis. Nat Cell Biol 8: 227–237.
Zhang H, Macara IG. 2008. The PAR-6 polarity protein regulates dendritic spine morphogenesis
through p190 RhoGAP and the Rho GTPase. Dev Cell 14: 216–226.
Zhang X, Zhu J, Yang GY, Wang QJ, Qian L, Chen YM, Chen F, Tao Y, Hu HS, Wang T, et al. 2007.
Dishevelled promotes axon differentiation by regulating atypical protein kinase C. Nat Cell Biol 9:
743–754.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a009654
CHAPTER 10
Correspondence: perrimon@receptor.med.harvard.edu
SUMMARY
Outline
1 Introduction
2 Embryonic patterning: Interplay between transcriptional cascades and
signaling
3 Juxtacrine signaling: Notch as an example
4 Patterning by secreted paracrine factors
5 Controlling the signaling range of secreted factors
6 The logic of signaling
7 Integrating signaling pathways
8 Concluding remarks: Developmental versus physiological signaling
References
1 INTRODUCTION
Key to multicellularity is the coordinated interaction of the various cells that
make up the body. Indeed, patterning of embryos, establishment of cell type
diversity, and formation of tissues and organs all rely on cell-to-cell
communication during development. Thus, arguably one of the most
important principles of developmental biology involves “one group of cells
changing the behavior of an adjacent set of cells, causing them to change
their shape, mitotic rate, or fate” (Gilbert 2000).
Classically, the ability of one group of cells to affect the fate of another is
called “induction.” The cells that produce the signals are referred to as
“inducing cells,” whereas the receiving cells are termed “responders”
(Spemann and Mangold 1924). The ability of cells to respond to the inducers,
referred to as “competence” (Waddington 1940), usually reflects the presence
of a receptor at the top of a pathway that regulates the expression of specific
transcription factors in the responding cells. The responding cells, in turn, can
become inductive and change the fate of their neighbors by producing new
signals, thus generating sequential inductive events that increase cell-fate
diversity in tissues.
Identification and characterization of the signaling pathways involved in
development has led to the surprising realization that only a few exist
(Gerhart 1999; Gilbert 2000; Barolo and Posakony 2002). These fall into 11
main classes, defined by the ligand or signal transducers involved: Notch,
FGF, EGF, Wnt/Wingless (Wg), Hedgehog (Hh), transforming growth factor
β (TGFβ)/BMPs, cytokine (nonreceptor tyrosine kinase JAK-STAT [signal
transducers and activators of transcription] pathway), Hippo, Jun kinase
(JNK), NF-κB, and retinoic acid receptor (RAR). These pathways involve
either cell-to-cell contact via surface proteins (juxtacrine signaling), or
secreted diffusible growth and differentiation factors (paracrine signaling).
Among the pathways mentioned above, only two of them, Notch and Hippo,
are juxtacrine, whereas the others are paracrine.
With the exception of those that release steroid hormones and retinoic
acid, which cross the membrane and activate gene expression by binding
directly to receptor proteins that act as transcriptional regulators, inducing
cells generally produce secreted or transmembrane ligands, which in some
cases require complex processing in the producing cells or the extracellular
matrix. When these ligands bind to transmembrane receptors on target cells
they activate a cascade of events that ultimately regulate the activity of a
small number of transcription factors and/or cofactors, triggering gene-
expression programs that drive the cellular changes. For example, Notch
signaling (p. 109 [Kopan 2012]) regulates CSL (for CBF1, Suppressor of
Hairless, and Lag1) proteins that possess an integrase domain, receptor
tyrosine kinases (RTKs) regulate ETS (erythroblast transformation-specific)
transcription factors, Wnt ligands (p. 103 [Nusse 2012]) mostly regulate the
high-mobility group (HMG) box-containing TCF (T-cell factor) transcription
factor, Hh proteins (p. 107 [Ingham 2012]) regulate Gli (glioblastoma)
transcription factors that have DNA-binding zinc-finger domains, and BMPs
(p. 113 [Wrana 2013]) regulate Smads (Sma- and Mad-related proteins)
transcription factors. Cytokine pathways (p. 117 [Harrison 2012]) regulate
STATs, and Hippo (p. 133 [Harvey and Hariharan 2012]) regulates TAZ (for
transcriptional coactivator with PDZ-binding motif) proteins that contain a
WW domain and a carboxy-terminal PDZ-binding motif (Table 1). In
addition, many pathways activate feedback loops that modulate or terminate
the incoming signal (Perrimon and McMahon 1999; Freeman 2000).
Figure 1. Patterning of the early Drosophila embryo. (A) Anterior–posterior patterning and
segmentation of the embryo is initiated by maternally deposited gene products that regulate the
expression of gap genes. Gap genes in turn control the expression of pair-rule genes, which themselves
regulate segment-polarity genes. The gene hierarchy and activation/repression interactions between
different transcription factors that coordinate patterning of the anterior–posterior axis of the early
Drosophila embryo are shown to the right. (B) Dorsal–ventral patterning is initiated by a Dorsal nuclear
gradient regulated by the Toll/NF-κB pathway. Graded nuclear localization of Dorsal subdivides the
dorso–ventral axis into distinct domains expressing twist (twi) and snail (sna), rhomboid (rho) and
decapendaplegic (dpp), zerknullt (zen) and tolloid (tld), which will form the prospective mesoderm,
neurogenic ectoderm, and dorsal ectoderm, respectively. Dorsal activates the zygotic transcription
program in the dorsoventral axis. (C) Terminal patterning is initiated in the germline by the localized
expression of Torsolike in the space outside the poles of the embryo. Torsolike activates the Torso
ligand (Trunk) locally and this is followed by Torso activation at the poles of the embryo, which will
lead to induction of the terminal patterning genes tailless and huckebein. Torso is an RTK and its action
is propagated through the MAPK pathway.
5.2 Hedgehog
HH transmits information over several cell diameters, but its range is
restricted. The distribution of HH has been studied most intensively in the
wing imaginal disc, where it defines a zone of activation in the boundary
between the posterior and anterior compartments of the disc. All posterior
cells produce HH but do not respond to it, whereas the anterior cells do not
produce it but can respond to it (review by Ingham and McMahon 2001). The
range of HH diffusion from the posterior compartment determines the
signaling range, and the region of the anterior compartment that receives HH
subsequently becomes the domain that produces the BMP family ligand DPP,
which directs long-range patterning of the wing.
HH is unusual as it undergoes dual lipid modification and autoproteolytic
cleavage (Porter et al. 1996; Pepinsky et al. 1998; Chen et al. 2004). The
cholesterol moiety that is added limits HH trafficking within and between
cells and palmitoylation is required for the production of a soluble multimeric
HH protein. Binding of HH to its receptor Patched (PTC) leads to its
endocytosis and degradation. Because PTC functions by inhibiting the next
step in the pathway (the transmembrane protein Smoothened [SMO]), this
leads to pathway activation. Interestingly, ptc itself is a transcriptional target
gene for HH signaling (reviewed in Wilson and Chuang 2010). As in the case
of Branchless, this leads to more effective trapping of HH by the first rows of
cells receiving the signal, and hence to a restriction of the signaling range.
Recently, studies in mammalian cells have shown that mammalian Hh
signaling depends on the primary cilium, a small cellular projection found on
most vertebrate cells (Goetz et al. 2009). In particular, Smo proteins
participate in the transduction of Hh signals, moving into the cilium in
response to Hh ligand. Interestingly, the absence of cilia in Drosophila
suggests that a fundamental difference exists between the organization of the
Hh pathway between invertebrates and mammals.
FGF and Wnt signaling are regulated temporally and spatially. The
ligands are expressed in gradients in the precursor tissue of the segments
where they regulate the progressive maturation/differentiation of the tissue
and define the domain of the clock activities. The Wnt pathway, for example,
is activated in the PSM before segmentation, plays a role upstream of both
the Wnt and Notch oscillations, and is thought to entrain the Notch feedback
loop. As a result, the spatial and highly dynamic temporal regulations of
these signaling activities guarantee the robust segmentation patterning of the
vertebrate axis and are evolutionarily conserved in vertebrates (Dequeant and
Pourquie 2008). In addition, experiments in zebrafish have indicated that the
Notch pathway is required for synchrony of the oscillations at the cellular
level and the coordinated expression of the correct targets within neighboring
cells, because lack of Notch leads to a “salt and pepper” pattern of
oscillations (Fig. 5A) (Dequeant and Pourquie 2008).
Determination of mesodermal progenitors in the Drosophila embryo
(Carmena et al. 1998) exemplifies the complex interplay and integration of
signaling pathways at the promoter level (Fig. 6). Using the regulation of the
even-skipped (eve) promoter, Halfon et al. (2000) have illustrated how the
synergistic integration of transcription factors, regulated by the Wnt/Wg,
DPP/BMP, and EGF/FGF/ERK pathways, generates a specific developmental
transcriptional response at a single defined enhancer. Because some of the
pathways are activated earlier than others and in a broader domain, they
determine the “competence group” of cells (expressing markers like Lethal of
scute, L’sc) and lead to subsequent activation of additional pathways within a
more restricted cell population (Fig. 6A). These later pathways are regarded
as inductive, and it is the final integration of the transcriptional signals from
all pathways, within a single enhancer, that induces the relevant target gene.
In this system, the WG and DPP signals are orthogonal to each other and
define the intersection zone as the competence group, through signal-
responsive transcription factors (MAD and TCF) that induce two tissue-
specific transcription factors (Tinman and Twist). In addition, TCF also
contributes to the expression of essential elements for ERK signaling (i.e.,
Rhomboid, Heartless, and Heartbroken). Once activated, ERK provides the
inductive signal, by activating the transcription factor Pointed and
inactivating the YAN repressor (Fig. 6B). Finally, singling out of
mesodermal Eve progenitors is achieved through the process of lateral
inhibition mediated by Notch/Delta signaling (Carmena et al. 2002).
Figure 6. The WG/DPP/FGF interplay during specification of Drosophila mesodermal progenitors. (A)
A model for patterning of the embryonic Drosophila mesoderm through the combinatorial actions of
WG, DPP, and RAS/ERK signals. This model applies to both somatic muscle and pericardial muscle.
The intersection between WG (red) and DPP (blue) delineates a prepattern (purple) in which Lethal of
scute (L’SC) is initially activated in a precluster (orange). The entire L’SC precluster is competent to
respond to RAS1. However, the spatially restricted activation of Heartless (HTL) and EGFR restricts
L’SC to a subset of precluster cells that correspond to an equivalence group. RAS1 signaling activates
EVE expression in all cells of the L’sc cluster (green) and subsequently a single EVE-expressing
progenitor (red) is determined by lateral inhibition mediated by the Notch/Delta pathway. (B) WG,
DPP, and RAS1 signal integration during specification of mesodermal EVE progenitors. WG and DPP
provide developmental competence by regulating tissue-specific transcription factors (Tinman [TIN]
and Twist [TWI]), signal-responsive transcription factors (MAD, TCF), and proximal components of
the RTK/ERK pathway (FGFR/HTL, Heartbroken [HBR]/DOF and Rhomboid [RHO]). The RAS
pathway leads to activation of the ETS-binding transcription factor Pointed (PNT) and inactivation of
the ETS-binding YAN repressor. The activities of all five transcriptional activators (TIN, TWI, MAD,
TCF, and PNT) are integrated at the MHE (Muscle and Heart Enhancer) of eve, which is located 6 kb
downstream from its transcription start site, and synergistically promote eve expression. In the absence
of inductive RAS signaling, YAN represses eve by binding to ETS sites. In addition, RAS/PNT
signaling in the EVE progenitor promotes Delta and Argos (AOS) expression, which in combination
activate Notch and shut down EGFR signaling in the nonprogenitor cells, to ensure lateral inhibition.
ACKNOWLEDGMENTS
We thank Mary-Lee Dequeant for helpful comments. Work in the Perrimon
laboratory is supported by the Howard Hughes Medical Institute and NIH,
and work in the Shilo laboratory by a grant from the Minerva Foundation.
REFERENCES
*Reference is in this book.
Affolter M, Basler K. 2007. The decapentaplegic morphogen gradient: From pattern formation to
growth regulation. Nat Rev Genet 8: 663–674.
Arbouzova NI, Zeidler MP. 2006. JAK/STAT signalling in Drosophila: Insights into conserved
regulatory and cellular functions. Development 133: 2605–2616.
Artavanis-Tsakonas S, Rand MD, Lake RJ. 1999. Notch signaling: Cell fate control and signal
integration in development. Science 284: 770–776.
Barolo S, Posakony JW. 2002. Three habits of highly effective signaling pathways: Principles of
transcriptional control by developmental cell signaling. Genes Dev 16: 1167–1181.
Bate M, Rushton E. 1993. Myogenesis and muscle patterning in Drosophila. C R Acad Sci III 316:
1047–1061.
Bray SJ. 2006. Notch signalling: A simple pathway becomes complex. Nat Rev Mol Cell Biol 7: 678–
689.
Carmena A, Gisselbrecht S, Harrison J, Jimenez F, Michelson AM. 1998. Combinatorial signaling
codes for the progressive determination of cell fates in the Drosophila embryonic mesoderm. Genes
Dev 12: 3910–3922.
Carmena A, Buff E, Halfon MS, Gisselbrecht S, Jimenez F, Baylies MK, Michelson AM. 2002.
Reciprocal regulatory interactions between the Notch and Ras signaling pathways in the Drosophila
embryonic mesoderm. Dev Biol 244: 226–242.
Casci T, Vinos J, Freeman M. 1999. Sprouty, an intracellular inhibitor of Ras signaling. Cell 96: 655–
665.
Chen MH, Li YJ, Kawakami T, Xu SM, Chuang PT. 2004. Palmitoylation is required for the
production of a soluble multimeric Hedgehog protein complex and long-range signaling in
vertebrates. Genes Dev 18: 641–659.
Cohen M, Georgiou M, Stevenson NL, Miodownik M, Baum B. 2010. Dynamic filopodia transmit
intermittent Delta-Notch signaling to drive pattern refinement during lateral inhibition. Dev Cell 19:
78–89.
De Joussineau C, Soule J, Martin M, Anguille C, Montcourrier P, Alexandre D. 2003. Delta-promoted
filopodia mediate long-range lateral inhibition in Drosophila. Nature 426: 555–559.
Dequeant ML, Pourquie O. 2008. Segmental patterning of the vertebrate embryonic axis. Nat Rev
Genet 9: 370–382.
Dequeant ML, Glynn E, Gaudenz K, Wahl M, Chen J, Mushegian A, Pourquie O. 2006. A complex
oscillating network of signaling genes underlies the mouse segmentation clock. Science 314: 1595–
1598.
Doroquez DB, Rebay I. 2006. Signal integration during development: Mechanisms of EGFR and Notch
pathway function and cross-talk. Crit Rev Biochem Mol Biol 41: 339–385.
Duffy JB, Perrimon N. 1994. The torso pathway in Drosophila: Lessons on receptor tyrosine kinase
signaling and pattern formation. Dev Biol 166: 380–395.
Edwin F, Anderson K, Ying C, Patel TB. 2009. Intermolecular interactions of Sprouty proteins and
their implications in development and disease. Mol Pharmacol 76: 679–691.
Eldar A, Rosin D, Shilo BZ, Barkai N. 2003. Self-enhanced ligand degradation underlies robustness of
morphogen gradients. Dev Cell 5: 635–646.
Feng XH, Derynck R. 2005. Specificity and versatility in tgf-β signaling through Smads. Annu Rev Cell
Dev Biol 21: 659–693.
Fior R, Henrique D. 2009. “Notch-Off”: A perspective on the termination of Notch signalling. Int J Dev
Biol 53: 1379–1384.
Flores GV, Duan H, Yan H, Nagaraj R, Fu W, Zou Y, Noll M, Banerjee U. 2000. Combinatorial
signaling in the specification of unique cell fates. Cell 103: 75–85.
Fortini ME. 2009. Notch signaling: The core pathway and its posttranslational regulation. Dev Cell 16:
633–647.
Freeman M. 2000. Feedback control of intercellular signalling in development. Nature 408: 313–319.
Friedman A, Perrimon N. 2007. Genetic screening for signal transduction in the era of network biology.
Cell 128: 225–231.
Gabay L, Scholz H, Golembo M, Klaes A, Shilo BZ, Klambt C. 1996. EGF receptor signaling induces
pointed P1 transcription and inactivates Yan protein in the Drosophila embryonic ventral ectoderm.
Development 122: 3355–3362.
Gabay L, Seger R, Shilo BZ. 1997. MAP kinase in situ activation atlas during Drosophila
embryogenesis. Development 124: 3535–3541.
Galli LM, Barnes TL, Secrest SS, Kadowaki T, Burrus LW. 2007. Porcupine-mediated lipid-
modification regulates the activity and distribution of Wnt proteins in the chick neural tube.
Development 134: 3339–3348.
Gerhart J. 1999. 1998 Warkany lecture: Signaling pathways in development. Teratology 60: 226–239.
Ghabrial A, Luschnig S, Metzstein MM, Krasnow MA. 2003. Branching morphogenesis of the
Drosophila tracheal system. Annu Rev Cell Dev Biol 19: 623–647.
Ghiglione C, Carraway KLIII, Amundadottir LT, Boswell RE, Perrimon N, Duffy JB. 1999. The
transmembrane molecule kekkon 1 acts in a feedback loop to negatively regulate the activity of the
Drosophila EGF receptor during oogenesis. Cell 96: 847–856.
Gilbert SF. 2000. Developmental biology, 6th ed. Sinauer Associates, Sunderland, MA.
Goetz SC, Ocbina PJ, Anderson KV. 2009. The primary cilium as a Hedgehog signal transduction
machine. Methods Cell Biol 94: 199–222.
Golembo M, Raz E, Shilo BZ. 1996. The Drosophila embryonic midline is the site of Spitz processing,
and induces activation of the EGF receptor in the ventral ectoderm. Development 122: 3363–3370.
Greco V, Hannus M, Eaton S. 2001. Argosomes: A potential vehicle for the spread of morphogens
through epithelia. Cell 106: 633–645.
Greenwald I. 1998. LIN-12/Notch signaling: Lessons from worms and flies. Genes Dev 12: 1751–
1762.
Greenwald I, Rubin GM. 1992. Making a difference: The role of cell-cell interactions in establishing
separate identities for equivalent cells. Cell 68: 271–281.
Hacohen N, Kramer S, Sutherland D, Hiromi Y, Krasnow MA. 1998. Sprouty encodes a novel
antagonist of FGF signaling that patterns apical branching of the Drosophila airways. Cell 92: 253–
263.
Halfon MS, Carmena A, Gisselbrecht S, Sackerson CM, Jimenez F, Baylies MK, Michelson AM. 2000.
Ras pathway specificity is determined by the integration of multiple signal-activated and tissue-
restricted transcription factors. Cell 103: 63–74.
* Harrison DA. 2012. The JAK/STAT pathway. Cold Spring Harb Perspect Biol 4: a011205.
* Harvey KF, Hariharan IK. 2012. The Hippo pathway. Cold Spring Harb Perspect Biol 4: a011288.
Heemskerk J, DiNardo S, Kostriken R, O’Farrell PH. 1991. Multiple modes of engrailed regulation in
the progression towards cell fate determination. Nature 352: 404–410.
Heitzler P, Simpson P. 1993. Altered epidermal growth factor-like sequences provide evidence for a
role of Notch as a receptor in cell fate decisions. Development 117: 1113–1123.
Hou SX, Zheng Z, Chen X, Perrimon N. 2002. The Jak/STAT pathway in model organisms: Emerging
roles in cell movement. Dev Cell 3: 765–778.
Huppert SS, Jacobsen TL, Muskavitch MA. 1997. Feedback regulation is central to Δ-Notch signalling
required for Drosophila wing vein morphogenesis. Development 124: 3283–3291.
Ikeda S, Kishida S, Yamamoto H, Murai H, Koyama S, Kikuchi A. 1998. Axin, a negative regulator of
the Wnt signaling pathway, forms a complex with GSK-3β and β-catenin and promotes GSK-3β-
dependent phosphorylation of β-catenin. EMBO J 17: 1371–1384.
* Ingham PW. 2012. Hedgehog signaling. Cold Spring Harb Perspect Biol 4: a011221.
Ingham PW, McMahon AP. 2001. Hedgehog signaling in animal development: Paradigms and
principles. Genes Dev 15: 3059–3087.
Jan YN, Jan LY. 1995. Maggot’s hair and bug’s eye: Role of cell interactions and intrinsic factors in
cell fate specification. Neuron 14: 1–5.
Jiang J, Hui CC. 2008. Hedgehog signaling in development and cancer. Dev Cell 15: 801–812.
Jung SH, Evans CJ, Uemura C, Banerjee U. 2005. The Drosophila lymph gland as a developmental
model of hematopoiesis. Development 132: 2521–2533.
Kadowaki T, Wilder E, Klingensmith J, Zachary K, Perrimon N. 1996. The segment polarity gene
porcupine encodes a putative multitransmembrane protein involved in Wingless processing. Genes
Dev 10: 3116–3128.
Klambt C. 1993. The Drosophila gene pointed encodes two ETS-like proteins which are involved in
the development of the midline glial cells. Development 117: 163–176.
Klein DE, Nappi VM, Reeves GT, Shvartsman SY, Lemmon MA. 2004. Argos inhibits epidermal
growth factor receptor signalling by ligand sequestration. Nature 430: 1040–1044.
* Kopan R. 2012. Notch signaling. Cold Spring Harb Perspect Biol 4: a011213.
Kornberg TB, Guha A. 2007. Understanding morphogen gradients: A problem of dispersion and
containment. Curr Opin Genet Dev 17: 264–271.
Kramer S, Okabe M, Hacohen N, Krasnow MA, Hiromi Y. 1999. Sprouty: A common antagonist of
FGF and EGF signaling pathways in Drosophila. Development 126: 2515–2525.
Lai EC. 2004. Notch signaling: Control of cell communication and cell fate. Development 131: 965–
973.
Lecuit T, Brook WJ, Ng M, Calleja M, Sun H, Cohen SM. 1996. Two distinct mechanisms for long-
range patterning by Decapentaplegic in the Drosophila wing. Nature 381: 387–393.
Lee JR, Urban S, Garvey CF, Freeman M. 2001. Regulated intracellular ligand transport and
proteolysis control EGF signal activation in Drosophila. Cell 107: 161–171.
Levine M. 2008. Dorsal-ventral patterning of the Drosophila embryo. In The legacy of Drosophila
genetics: From “defining the gene” to “analyzing genome function” (ed. Bier E). Henry Stewart
Talks, London.
Logan CY, Nusse R. 2004. The Wnt signaling pathway in development and disease. Annu Rev Cell Dev
Biol 20: 781–810.
MacDonald BT, Tamai K, He X. 2009. Wnt/β-catenin signaling: Components, mechanisms, and
diseases. Dev Cell 17: 9–26.
Meinhardt H. 1978. Space-dependent cell determination under the control of morphogen gradient. J
Theor Biol 74: 307–321.
Melen GJ, Levy S, Barkai N, Shilo BZ. 2005. Threshold responses to morphogen gradients by zero-
order ultrasensitivity. Mol Syst Biol 1: 2005.0028.
Nagaraj R, Banerjee U. 2004. The little R cell that could. Int J Dev Biol 48: 755–760.
Nellen D, Burke R, Struhl G, Basler K. 1996. Direct and long-range action of a DPP morphogen
gradient. Cell 85: 357–368.
Neumann CJ, Cohen SM. 1997. Long-range action of Wingless organizes the dorsal-ventral axis of the
Drosophila wing. Development 124: 871–880.
Noselli S, Perrimon N. 2000. Signal transduction. Are there close encounters between signaling
pathways? Science 290: 68–69.
* Nusse R. 2012. Wnt signaling. Cold Spring Harb Perspect Biol 4: a011163.
Nusslein-Volhard C, Wieschaus E. 1980. Mutations affecting segment number and polarity in
Drosophila. Nature 287: 795–801.
O’Neill EM, Rebay I, Tjian R, Rubin GM. 1994. The activities of two Ets-related transcription factors
required for Drosophila eye development are modulated by the Ras/MAPK pathway. Cell 78: 137–
147.
Ohshiro T, Emori Y, Saigo K. 2002. Ligand-dependent activation of breathless FGF receptor gene in
Drosophila developing trachea. Mech Dev 114: 3–11.
Pan D. 2007. Hippo signaling in organ size control. Genes Dev 21: 886–897.
Panakova D, Sprong H, Marois E, Thiele C, Eaton S. 2005. Lipoprotein particles are required for
Hedgehog and Wingless signalling. Nature 435: 58–65.
Panin VM, Papayannopoulos V, Wilson R, Irvine KD. 1997. Fringe modulates Notch-ligand
interactions. Nature 387: 908–912.
Pepinsky RB, Zeng C, Wen D, Rayhorn P, Baker DP, Williams KP, Bixler SA, Ambrose CM, Garber
EA, Miatkowski K, et al. 1998. Identification of a palmitic acid-modified form of human Sonic
hedgehog. J Biol Chem 273: 14037–14045.
Perrimon N, Bernfield M. 2000. Specificities of heparan sulphate proteoglycans in developmental
processes. Nature 404: 725–728.
Perrimon N, McMahon AP. 1999. Negative feedback mechanisms and their roles during pattern
formation. Cell 97: 13–16.
Perrimon N, Perkins L. 1997. There must be 50 ways to rule the signal: The case of the Drosophila
EGF receptor. Cell 89: 13–16.
Porter JA, Ekker SC, Park WJ, von Kessler DP, Young KE, Chen CH, Ma Y, Woods AS, Cotter RJ,
Koonin EV, et al. 1996. Hedgehog patterning activity: Role of a lipophilic modification mediated
by the carboxy-terminal autoprocessing domain. Cell 86: 21–34.
Pownall ME, Isaacs HV. 2010. FGF signalling in vertebrate development. Morgan & Claypool Life
Sciences, San Rafael, CA.
Rebay I, Rubin GM. 1995. Yan functions as a general inhibitor of differentiation and is negatively
regulated by activation of the Ras1/MAPK pathway. Cell 81: 857–866.
Reddi HV, Madde P, Marlow LA, Copland JA, McIver B, Grebe SK, Eberhardt NL. 2010. Expression
of the PAX8/PPARγ fusion protein is associated with decreased neovascularization in vivo: Impact
on tumorigenesis and disease prognosis. Genes Cancer 1: 480–492.
Reich A, Sapir A, Shilo B. 1999. Sprouty is a general inhibitor of receptor tyrosine kinase signaling.
Development 126: 4139–4147.
Roignant JY, Treisman JE. 2009. Pattern formation in the Drosophila eye disc. Int J Dev Biol 53: 795–
804.
Roy S, Hsiung F, Kornberg TB. 2011. Specificity of Drosophila cytonemes for distinct signaling
pathways. Science 332: 354–358.
Rusch J, Levine M. 1996. Threshold responses to the dorsal regulatory gradient and the subdivision of
primary tissue territories in the Drosophila embryo. Curr Opin Genet Dev 6: 416–423.
Rushton E, Drysdale R, Abmayr SM, Michelson AM, Bate M. 1995. Mutations in a novel gene,
myoblast city, provide evidence in support of the founder cell hypothesis for Drosophila muscle
development. Development 121: 1979–1988.
Sandmann T, Girardot C, Brehme M, Tongprasit W, Stolc V, Furlong EE. 2007. A core transcriptional
network for early mesoderm development in Drosophila melanogaster. Genes Dev 21: 436–449.
Saucedo LJ, Edgar BA. 2007. Filling out the Hippo pathway. Nat Rev Mol Cell Biol 8: 613–621.
Shilo BZ. 2005. Regulating the dynamics of EGF receptor signaling in space and time. Development
132: 4017–4027.
Shilo BZ, Schejter ED. 2011. Regulation of developmental intercellular signalling by intracellular
trafficking. EMBO J 30: 3516–3526.
Smith WC. 1999. TGFβ inhibitors. New and unexpected requirements in vertebrate development.
Trends Genet 15: 3–5.
Spemann H, Mangold H. 1924. Über induktion von Embryonalagen durch Implantation Artfremder
Organisatoren. Roux’ Arch Entw Mech 100: 599–638.
St Johnston D, Nusslein-Volhard C. 1992. The origin of pattern and polarity in the Drosophila embryo.
Cell 68: 201–219.
Strigini M, Cohen SM. 2000. Wingless gradient formation in the Drosophila wing. Curr Biol 10: 293–
300.
Tsruya R, Schlesinger A, Reich A, Gabay L, Sapir A, Shilo BZ. 2002. Intracellular trafficking by Star
regulates cleavage of the Drosophila EGF receptor ligand Spitz. Genes Dev 16: 222–234.
Tsruya R, Wojtalla A, Carmon S, Yogev S, Reich A, Bibi E, Merdes G, Schejter E, Shilo BZ. 2007.
Rhomboid cleaves Star to regulate the levels of secreted Spitz. EMBO J 26: 1211–1220.
Tsuda L, Nagaraj R, Zipursky SL, Banerjee U. 2002. An EGFR/Ebi/Sno pathway promotes δ
expression by inactivating Su(H)/SMRTER repression during inductive notch signaling. Cell 110:
625–637.
Tsuneizumi K, Nakayama T, Kamoshida Y, Kornberg TB, Christian JL, Tabata T. 1997. Daughters
against dpp modulates dpp organizing activity in Drosophila wing development. Nature 389: 627–
631.
Turing AM. 1952. The chemical basis of morphogenesis. Philos Trans R Soc London B 237: 37–72.
Urban S, Lee JR, Freeman M. 2001. Drosophila rhomboid-1 defines a family of putative
intramembrane serine proteases. Cell 107: 173–182.
Varelas X, Miller BW, Sopko R, Song S, Gregorieff A, Fellouse FA, Sakuma R, Pawson T, Hunziker
W, McNeill H, et al. 2010. The Hippo pathway regulates Wnt/β-catenin signaling. Dev Cell 18:
579–591.
Waddington CH. 1940. Organizers and genes. Cambridge University Press, Cambridge.
Wartlick O, Kicheva A, Gonzalez-Gaitan M. 2009. Morphogen gradient formation. Cold Spring Harb
Perspect Biol 1: a001255.
Wilson CW, Chuang PT. 2010. Mechanism and evolution of cytosolic Hedgehog signal transduction.
Development 137: 2079–2094.
Wolpert L. 1969. Positional information and the spatial pattern of cellular differentiation. J Theor Biol
25: 1–47.
* Wrana JL. 2013. Signaling by the TGFβ superfamily. Cold Spring Harb Perspect Biol 5: a011197.
Wu MY, Hill CS. 2009. Tgf-β superfamily signaling in embryonic development and homeostasis. Dev
Cell 16: 329–343.
Xia L, Jia S, Huang S, Wang H, Zhu Y, Mu Y, Kan L, Zheng W, Wu D, Li X, et al. 2010. The
Fused/Smurf complex controls the fate of Drosophila germline stem cells by generating a gradient
BMP response. Cell 143: 978–990.
Yan D, Lin X. 2009. Shaping morphogen gradients by proteoglycans. Cold Spring Harb Perspect Biol
1: a002493.
Yogev S, Schejter ED, Shilo BZ. 2008. Drosophila EGFR signalling is modulated by differential
compartmentalization of Rhomboid intramembrane proteases. EMBO J 27: 1219–1230.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a005975
CHAPTER 11
SUMMARY
Outline
1 Introduction
2 Receptors: Detection and transduction
3 Structural basis of sensory receptor activation
4 The logic of sensory coding
5 Evolution and variation
6 Concluding remarks
References
1 INTRODUCTION
An organism’s perception of the world is filtered through its sensory systems.
The properties of these systems dictate the types of stimuli that can be
detected and constrain the ways in which these stimuli are reconstructed,
integrated, and interpreted. Here we discuss how sensory signals are received
and transduced, focusing on the first steps in the complex process of
perceiving an external stimulus. A recurrent theme is the way in which the
biochemical and biophysical properties of sensory receptor molecules, and
the neurons in which they reside, have been sculpted by evolution to capture
those signals that are most salient for the survival and reproduction of the
organism. As a result, some classes of sensory receptors, such as the night
vision receptor rhodopsin, show great conservation, whereas others, such as
olfactory receptors, show great diversity.
Evolutionary comparisons are fascinating at many levels, not least of
which is their power to highlight the logic of the stimulus–response
relationship. For example, honeybees can see UV light, enabling them to
locate sources of nectar and pollen based on the UV reflectance of flower
petals (Kevan et al. 2001), whereas humans and Old World primates have
excellent sensitivity and chromatic discrimination at longer wavelengths,
permitting the identification of red, orange, and yellow fruit against a
background of green foliage (Mollon 1989). Star-nosed moles use a
specialized mechanoreceptive organ on their snout to locate meals and
navigate through lightless subterranean tunnels (Catania 2005), and pit vipers
have evolved thermoreceptive organs to detect the infrared radiation emitted
by their warm-blooded prey (Campbell et al. 2002). In each of these cases,
evolution has fine-tuned a sensory organ through anatomical and/or
molecular changes to enhance the detection of relevant stimuli.
For simplicity, we focus on eukaryotic sensory systems, in which G-
protein-coupled receptors (GPCRs) and ion channels predominate as sensory
receptors. The one exception is mechanosensation, in which the molecular
basis of membrane stretch detection has been beautifully delineated in
bacteria but remains less clear in eukaryotes. Thus, our discussion of
mechanosensation is focused largely on prokaryotic systems. We also
describe the diverse cast of downstream transduction pathways and the
manner in which receptors and transduction pathways are regulated to
terminate signaling and set receptor sensitivity.
Before discussing individual receptors, it is worth noting that the
physiological attributes of sensory systems are dictated not only by the
molecular properties of receptor molecules and their associated signal
transduction proteins, but also by the architecture of the sensory organs, cells,
and subcellular structures in which they reside. A notable example is the
retina, where visual pigments, together with other components of the
phototransduction pathway, are localized within outer segments of rod and
cone photoreceptor cells at near millimolar concentrations, thereby enhancing
light sensitivity and transduction efficiency (Yau and Hardie 2009). Another
example is seen in the cochlea, where sound pressure waves are transmitted
through the middle ear to induce a localized and frequency-dependent
distortion of the basilar membrane in the cochlea. Progressive changes in
both the mechanical properties of the basilar membrane and the electrical
properties of the auditory hair cells along the length of the cochlea generate a
tonotopic map in which the amplitudes of different frequency components in
a complex sound are reflected in the magnitudes of auditory receptor
activation at different locations within the cochlea (Roberts et al. 1988).
Figure 5. Chromaticity diagrams for humans and honeybees. The triangles represent a plane within a
3D receptor space, with each vertex corresponding to the point at which the plane intersects an axis
representing the degree of excitation of one receptor type. (S) Short-wavelength receptor axis; (M)
medium-wavelength receptor axis; (L) long-wavelength receptor axis. The small black circles within
the triangles represent the chromaticities of a set of fruits that are consumed by primates (upper panels)
or a set of flower petals (lower panels). The line within each chromaticity diagrams represents the locus
of spectrally pure lights, with black circles and the adjacent numbers marking steps of 50 nm. (Adapted
from Osorio and Vorobyev 2008; reprinted, with permission, from Elsevier © 2008.)
6 CONCLUDING REMARKS
The sensory receptors and signaling systems described here represent only a
small sampling of the many that have been investigated. Because of space
limitations, we have not discussed plant sensory systems, and we have only
briefly touched on microbial systems. Even with this limited sampling, the
diversity of sensory systems is striking, and it is evident that each lifestyle
and ecological niche is accompanied by a distinctive set of adaptations in
sensory system structure and function.
Over the past 30 years, the identities and primary structures of most of the
major classes of vertebrate sensory receptor proteins have been defined. The
one exception is the mechanosensory channels in the auditory and vestibular
system, which remain enigmatic. The signaling cascades downstream from
GPCR-type sensory receptors have also been largely defined. In vertebrate
photoreception, the best studied of all GPCR signaling cascades, it is likely
that all of the components involved in signaling and adaptation are now
known. In less experimentally accessible systems, such as the taste and
vomeronasal systems, additional signaling components remain to be
identified, and the interactions between signal activation pathways and
feedback loops are still incompletely understood.
A full molecular understanding of sensory receptor function requires the
3D structures of receptors and signaling components in their various active
and inactive conformations, and, in some cases, in complex with each other.
This has been achieved for rhodopsin, transducin, and the bacterial MscL
mechanosensory channel, and it is an area of active investigation for other
classes of sensory receptors and their downstream effectors. Structural
studies can be expected to play a critical role in elucidating the molecular
basis of receptor–ligand specificity in chemosensory systems.
A further challenge in the field of sensory biology comes from the need to
diagnose and treat diseases that affect sensory signaling, including those
associated with chronic pain or the loss of vision or hearing. Understanding
sensory signaling at the molecular and cellular levels will inform these
clinical investigations and will continue to be one of nature’s grand scientific
challenges for biologists, chemists, physicists, and engineers.
REFERENCES
Abrahamson EW, Japar SM. 1972. Principles of the interaction of light and matter. In Handbook of
sensory physiology. VII/1: Photochemistry of vision (ed. Dartnall HJA), pp. 1–32. Springer-Verlag,
Berlin.
Arshavsky VY, Lamb TD, Pugh EN. 2002. G proteins and phototransduction. Annu Rev Physiol 64:
153–187.
Bakken GS, Krochmal AR. 2007. The imaging properties and sensitivity of the facial pit of pitvipers as
determined by optical and heat-transfer analysis. J Exp Biol 210: 2801–2810.
Basbaum AI, Bautista DM, Scherrer G, Julius D. 2009. Cellular and molecular mechanisms of pain.
Cell 139: 267–284.
Baylor DA, Matthews G, Yau KW. 1980. Two components of electrical dark noise in toad retinal rod
outer segments. J Physiol 309: 591–621.
Benton R, Vannice KS, Gomez-Diaz C, Vosshall LB. 2009. Variant ionotropic glutamate receptors as
chemosensory receptors in Drosophila. Cell 136: 149–162.
Billig GM, Pál B, Fidzinski P, Jentsch TJ. 2011. Ca2+-activated Cl− currents are dispensable for
olfaction. Nat Neurosci 14: 763–769.
Campbell AL, Naik RR, Sowards L, Stone MO. 2002. Biological infrared imaging and sensing. Micron
33: 211–225.
Carey AF, Wang G, Su CY, Zwiebel LJ, Carlson JR. 2010. Odorant reception in the malaria mosquito
Anopheles gambiae. Nature 464: 66–71.
Catania KC. 2005. Star-nosed moles. Curr Biol 15: R863–R864.
Caterina MJ, Julius D. 2001. The vanilloid receptor: A molecular gateway to the pain pathway. Annu
Rev Neurosci 24: 487–517.
Chalfie M. 2009. Neurosensory mechanotransduction. Nat Rev Mol Cell Biol 10: 44–52.
Chandrashekar J, Hoon MA, Ryba NJ, Zuker CS. 2006. The receptors and cells for mammalian taste.
Nature 444: 288–294.
Choe HW, Kim YJ, Park JH, Morizumi T, Pai EF, Krauss N, Hofmann KP, Scheerer P, Ernst OP.
2011. Crystal structure of metarhodopsin. II. Nature 471: 651–655.
DasGupta S, Wadell S. 2008. Learned odor discrimination in Drosophila without combinatorial odor
maps in the antennal lobe. Curr Biol 18: 1668–1674.
DeWire SM, Ahn S, Lefkowitz RJ, Shenoy SK. 2007. β-Arrestins and cell signaling. Annu Rev Physiol
69: 483–510.
Dohlman HG, Thorner J. 1997. RGS proteins and signaling by heterotrimeric G proteins. J Biol Chem
272: 3871–3874.
Dong D, Jones G, Zhang S. 2009. Dynamic evolution of bitter taste receptor genes in vertebrates. BMC
Evol Biol 9: 12.
Douglas RH, Partridge JC, Marshall NJ. 1998. The eyes of deep-sea fish. I: Lens pigmentation, tapeta
and visual pigments. Prog Retin Eye Res 17: 597–636.
Fasick JI, Cronin TW, Hunt DM, Robinson PR. 1998. The visual pigments of the bottlenose dolphin
(Tursiops truncatus). Vis Neurosci 15: 643–651.
Fasick JI, Robinson PR. 2000. Spectral-tuning mechanisms of marine mammal rhodopsins and
correlations with foraging depth. Vis Neurosci 17: 781–788.
Ferkey DM, Hyde R, Haspel G, Dionne HM, Hess HA, Suzuki H, Schafer WR, Koelle MR, Hart AC.
2007. C. elegans G protein regulator RGS-3 controls sensitivity to sensory stimuli. Neuron 53: 39–
52.
Freitag J, Ludwig G, Andreini I, Rössler P, Breer H. 1998. Olfactory receptors in aquatic and terrestrial
vertebrates. J Comp Physiol A 183: 635–650.
Gardiner JM, Atema J. 2010. The function of bilateral odor arrival time differences in olfactory
orientation of sharks. Curr Biol 20: 1187–1191.
Gaudet R. 2008. TRP channels entering the structural era. J Physiol 586: 3565–3575.
Gillespie PG, Walker RG. 2001. Molecular basis of mechanosensory transduction. Nature 413: 194–
202.
Gracheva EO, Ingolia NT, Kelly YM, Cordero-Morales JF, Hollopeter G, Chesler AT, Sánchez EE,
Perez JC, Weissman JS, Julius D. 2010. Molecular basis of infrared detection by snakes. Nature
464: 1006–1011.
Hallem EA, Carlson JR. 2006. Coding of odors by a receptor repertoire. Cell 125: 143–160.
Hallem EA, Ho MG, Carlson JR. 2004. The molecular basis of odor coding in the Drosophila antenna.
Cell 117: 965–979.
Hecht S, Schlaer S, Pirenne MH. 1942. Energy, quanta, and vision. J Gen Physiol 25: 819–840.
Henrik-Heldin C, Evans R, Gutkind S. 2012. Cold Spring Harb Perspect Biol (in press).
Iggo A. 1982. Cutaneous sensory mechanisms. In The senses (ed. Barlow HB, Mollon JD), pp. 369–
408. Cambridge University Press, Cambridge.
Jacobs GH, Neitz M, Neitz J. 1996. Mutations in S-cone pigment genes and the absence of colour
vision in two species of nocturnal primate. Proc Biol Sci 263: 705–710.
Johnstone BM, Patuzzi R, Yates GK. 1986. Basilar membrane measurements and the travelling wave.
Hear Res 22: 147–153.
Jones WD, Cayirlioglu P, Kadow IG, Vosshall LB. 2007. Two chemosensory receptors together
mediate carbon dioxide detection in Drosophila. Nature 445: 86–90.
Jordt SE, Julius D. 2002. Molecular basis for species-specific sensitivity to “hot” chili peppers. Cell
108: 421–430.
Kawamura S, Kubotera N. 2004. Ancestral loss of short wave-sensitive cone visual pigment in
lorisiform prosimians, contrasting with its strict conservation in other prosimians. J Mol Evol 58:
314–321.
Kevan PG, Chittka L, Dyer AG. 2001. Limits to the salience of ultraviolet: Lessons from colour vision
in bees and birds. J Exp Biol 204: 2571–2580.
Krispel CM, Chen D, Melling N, Chen YJ, Martemyanov KA, Quillinan N, Arshavsky VY, Wensel
TG, Chen CK, Burns ME. 2006. RGS expression rate-limits recovery of rod photoresponses.
Neuron 51: 409–416.
Kung C. 2005. A possible unifying principle for mechanosensation. Nature 436: 647–654.
Kung C, Martinac B, Sukharev S. 2010. Mechanosensitive channels in microbes. Annu Rev Microbiol
64: 313–329.
Li M, Yu Y, Yang J. 2011. Structural biology of TRP channels. Adv Exp Med Biol 704: 1–23.
Livingstone M, Hubel D. 1988. Segregation of form, color, movement, and depth: Anatomy,
physiology, and perception. Science 240: 740–749.
Lu T, Qiu YT, Wang G, Kwon JY, Rutzler M, Kwon HW, Pitts RJ, van Loon JJ, Takken W, Carlson
JR, et al. 2007. Odor coding in the maxillary palp of the malaria vector mosquito Anopheles
gambiae. Curr Biol 17: 1533–1544.
Lumpkin EA, Caterina MJ. 2007. Mechanisms of sensory transduction in the skin. Nature 445: 858–
865.
Lythgoe JN. 1979. The ecology of vision. Clarendon Press, Oxford.
Malnic B, Hirono J, Sato T, Buck LB. 1999. Combinatorial receptor codes for odors. Cell 96: 713–723.
Maloney LT. 1986. Evaluation of linear models of surface spectral reflectance with a small number of
parameters. J Opt Soc Am A 3: 1673–1683.
Milligan G. 2009. G protein-coupled receptor hetero-dimerization: Contribution to pharmacology and
function. Br J Pharmacol 58: 5–14.
Mollon JD. 1989. “Tho she kneel’d in that Place where they grew…”: The uses and origins of primate
colour vision. J Exp Biol 146: 21–38.
Moore CA, Milano SK, Benovic JL. 2007. Regulation of receptor trafficking by GRKs and arrestins.
Annu Rev Physiol 69: 451–482.
Myers BR, Sigal YM, Julius D. 2009. Evolution of thermal response properties in a cold-activated TRP
channel. PLoS One 4: e5741.
Neitzel KL, Hepler JR. 2006. Cellular mechanisms that determine selective RGS protein regulation of
G protein-coupled receptor signaling. Semin Cell Dev Biol 17: 383–389.
Niimura Y. 2009. On the origin and evolution of vertebrate olfactory receptor genes: Comparative
genome analysis among 23 chordate species. Genome Biol Evol 1: 34–44.
Niimura Y, Nei M. 2007. Extensive gains and losses of olfactory receptor genes in mammalian
evolution. PLoS One 2: e708.
Oka Y, Omura M, Kataoka H, Touhara K. 2004. Olfactory receptor antagonism between odorants.
EMBO J 23: 120–126.
Oka Y, Takai Y, Touhara K. 2009. Nasal airflow rate affects the sensitivity and pattern of glomerular
odorant responses in the mouse olfactory bulb. J Neurosci 29: 12070–12078.
Osorio D, Vorobyev M. 2008. A review of the evolution of animal colour vision and visual
communication signals. Vis Res 48: 2042–2051.
Perozo E, Rees DC. 2003. Structure and mechanism in prokaryotic mechanosensitive channels. Curr
Opin Struct Biol 13: 432–442.
Rasmussen SG, DeVree BT, Zou Y, Kruse AC, Chung KY, Kobilka TS, Thian FS, Chae PS, Pardon E,
Calinski D, et al. 2011. Crystal structure of the β2 adrenergic receptor-Gs protein complex. Nature
477: 549–555.
Regan BC, Julliot C, Simmen B, Viénot F, Charles-Dominique P, Mollon JD. 2001. Fruits, foliage and
the evolution of primate colour vision. Philos Trans R Soc Lond B Biol Sci 356: 229–283.
Roberts WM, Howard J, Hudspeth AJ. 1988. Hair cells: Transduction, tuning, and transmission in the
inner ear. Annu Rev Cell Biol 4: 63–92.
Rosenbaum DM, Rasmussen SG, Kobilka BK. 2009. The structure and function of G-protein-coupled
receptors. Nature 459: 356–363.
Ross EM, Wilkie TM. 2000. GTPase-activating proteins for heterotrimeric G proteins: Regulators of G
protein signaling (RGS) and RGS-like proteins. Annu Rev Biochem 69: 795–827.
Schnapf JL, Nunn BJ, Meister M, Baylor DA. 1990. Visual transduction in cones of the monkey
Macaca fascicularis. J Physiol 427: 681–713.
Schneeweis DM, Schnapf JL. 1999. The photovoltage of macaque cone photoreceptors: Adaptation,
noise, and kinetics. J Neurosci 19: 1203–1216.
Standfuss J, Edwards PC, D’Antona A, Fransen M, Xie G, Oprian DD, Schertler GF. 2011. The
structural basis of agonist-induced activation in constitutively active rhodopsin. Nature 471: 656–
660.
Stryer L. 1986. Cyclic GMP cascade of vision. Annu Rev Neurosci 9: 87–119.
Su CY, Menuz K, Carlson JR. 2009. Olfactory perception: Receptors, cells, and circuits. Cell 139: 45–
59.
Su CY, Martelli C, Emonet T, Carlson JR. 2011. Temporal coding of odor mixtures in an olfactory
receptor neuron. Proc Natl Acad Sci 108: 5075–5080.
Sukharev S. 2002. Purification of the small mechanosensitive channel of Escherichia coli (MscS): The
subunit structure, conduction, and gating characteristics in liposomes. Biophys J 83: 290–298.
Sukharev SI, Blount P, Martinac B, Blattner FR, Kung C. 1994. A large-conductance mechanosensitive
channel in E. coli encoded by mscL alone. Nature 368: 265–268.
Touhara K, Vosshall LB. 2009. Sensing odorants and pheromones with chemosensory receptors. Annu
Rev Physiol 71: 307–332.
Tricas TC, Kajiura SM, Summers AP. 2009. Response of the hammerhead shark olfactory epithelium
to amino acid stimuli. J Comp Physiol A Neuroethol Sens Neural Behav Physiol 195: 947–954.
Vásquez V, Sotomayor M, Cordero-Morales J, Schulten K, Perozo E. 2008. A structural mechanism for
MscS gating in lipid bilayers. Science 321: 1210–1214.
Vosshall LB. 2008. Scent of a fly. Neuron 59: 685–689.
Wang G, Carey AF, Carlson JR, Zwiebel LJ. 2010. Molecular basis of odor coding in the malaria
vector mosquito Anopheles gambiae. Proc Natl Acad Sci 107: 4418–4423.
Xu F, Wu H, Katritch V, Han GW, Jacobson KA, Gao ZG, Cherezov V, Stevens RC. 2011. Structure
of an agonist-bound human A2A adenosine receptor. Science 332: 322–327.
Yarmolinsky DA, Zuker CS, Ryba NJP. 2009. Common sense about taste: From mammals to insects.
Cell 139: 234–244.
Yau KW. 1991. Calcium and light adaptation in retinal photoreceptors. Curr Opin Neurobiol 1: 252–
257.
Yau KW, Hardie RC. 2009. Phototransduction motifs and variations. Cell 139: 246–264.
Young T. 1802. The Bakerian lecture. On the theory of light and colours. Phil Trans Roy Soc 92: 12–
48.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a005991
CHAPTER 12
Mary B. Kennedy
Division of Biology and Biological Engineering, California Institute of
Technology, Pasadena, California 91125
Correspondence: kennedym@its.caltech.edu
SUMMARY
1 INTRODUCTION
The function of the brain is to process and store information about the
environment and direct behavior in response to that information. Three major
cell types in the brain—excitatory neurons that use glutamate as their
transmitter, inhibitory neurons that use γ-aminobutyric acid (GABA) as their
transmitter, and glial cells—work together to respond to the environment
while maintaining the overall connectivity among neurons within an
acceptable homeostatic range. Excitatory neurons, the most numerous in the
brain, each receive thousands of synaptic inputs and, in turn, make thousands
of synaptic connections onto other neurons. A human brain contains, on
average, 86 billion neurons (Herculano-Houzel 2009) that in toto make
trillions of synaptic connections.
A typical cortical excitatory neuron (Fig. 1) comprises a neuronal soma
(cell body), several branched dendrites, and a single axon that can extend for
many millimeters and often branches to make thousands of individual
synaptic connections. The soma is the site of the nucleus and most of the
neuron’s protein synthetic machinery. Most inhibitory synaptic contacts
occur on the somal plasma membrane. In contrast, the highly branched
dendrites receive most of the excitatory synaptic contacts, which are made
onto small membrane protuberances called dendritic spines (Fig. 1). When
the neuronal membrane becomes depolarized to a threshold level, an action
potential is initiated at the base of the axon near the soma; this wave of
depolarization travels unabated to each of the thousands of presynaptic
endings along the axon. Depolarization causes membranous synaptic vesicles
within the presynaptic terminals to fuse with the plasma membrane at the
“active zone” opposite the postsynaptic site and flood the synaptic cleft with
neurotransmitters. Some of the transmitter molecules bind to specific
receptors, which are ligand-gated channels in the postsynaptic membrane.
Sodium and potassium ions flow through the channels of the activated
receptors, decreasing the gradient in their concentration across the membrane
and thus producing a localized depolarization called an excitatory
postsynaptic potential (EPSP). If the synapse fires repeatedly, or if several
different synapses on a neuron fire at the same time, the EPSPs can sum to
produce a depolarization that extends to the soma and initiates an action
potential.
Figure 1. Features of excitatory neurons in the brain. The cell body (soma) of a typical excitatory
pyramidal neuron is ∼10 μm in diameter and is located in one of several sheets of tightly packed somas
that define the layers of the neocortex and hippocampus. Apical and basal dendrites extend from the
soma, reaching into adjacent areas that are referred to as neuropil. Postsynaptic structures are located in
tiny membrane protuberances called spines that can be seen along the dendrites. Each soma gives rise
to one axon, which has a smaller diameter than the numerous dendrites. The axon can extend for
millimeters from the soma and branches to form thousands of presynaptic terminals where transmitter
is released onto the postsynaptic sites of other neurons. The axon hillock is located at the base of the
axon. Action potentials are usually initiated at this site; they travel along the axon (arrows) to
presynaptic terminals keeping a uniform amplitude of depolarization. Back-propagating action
potentials travel in the opposite direction through the soma and into the dendrites. The size of their
depolarization decreases as they travel and is regulated by the composition of dendritic ion channels.
(From Peters and Kaiserman-Abramof 1970; modified, with permission, © Wiley.)
2 SPINE SYNAPSES
In the brain, most synapses between excitatory neurons are located on spines,
tiny compartments that protrude from the neuron’s highly branched dendrites
(Fig. 2). A typical excitatory pyramidal neuron in the hippocampus or cortex
has ∼10,000 such synapses, most spines hosting just a single synapse. Spines
vary in size from ∼0.5–2 µm in length, from ∼0.25–1 µm in width, and from
∼10–100 attoliters in volume. The synaptic contact itself usually occurs at the
tip of the spine. It comprises the presynaptic active zone, the synaptic cleft,
and the postsynaptic receptor cluster and varies in diameter from ∼0.1 to 0.8
µm. Spines also vary in shape from stubby to thin to “mushroom-shaped.” In
general, the larger, mushroom-shaped spines contain stronger synapses.
Functionally, a stronger synapse is defined as one that contributes more
depolarization to the neuronal membrane upon activation than a weaker one;
thus, its activation is more likely to generate an action potential in the
postsynaptic neuron.
Figure 2. Synaptic plasticity. At the cellular level, one of the most essential elements of memory
formation is the adjustment in synaptic strength of excitatory synapses between neurons. AMPA-type
glutamate receptors (yellow) allow passage of sodium and potassium through their channel. Their
principal function is to depolarize the membrane, producing an excitatory postsynaptic potential
(EPSP). NMDA-type glutamate receptors (blue) also depolarize the membrane, but in addition to
sodium and potassium, calcium flows through their channel and can initiate synaptic plasticity. A long-
lasting increase in synaptic strength is referred to as long-term potentiation (LTP). LTP involves the
addition of new synaptic AMPA-type glutamate receptors (AMPARs) and an increase in the size of the
head of the postsynaptic spine, supported by an increase in the size and branching of the actin
cytoskeleton. Long-term depression (LTD) is a long-lasting decrease in synaptic strength that involves
a decrease in the number of synaptic AMPARs and shrinkage of the spine head. LTP is induced when
repeated firing of an action potential in the presynaptic terminal and the resulting release of glutamate
cause firing of action potentials in the postsynaptic neuron. LTD is induced when repeated firing of an
action potential in the presynaptic terminal does not cause firing of action potentials in the postsynaptic
neuron.
3.2.1 CaMKII
CaMKII makes up ∼1% of total protein in the forebrain and ∼2% in the
hippocampus (Bennett et al. 1983; Erondu and Kennedy 1985). These high
levels of expression, at least 10 times higher than those of other signaling
enzymes, are a specialization of excitatory neurons, which represent most of
the mass of the forebrain (Sik et al. 1998). CaMKII is present throughout the
cytosol of somas, axons, and dendrites, including spines in which it is present
both in the cytosol and the PSD (Kennedy et al. 1983; Chen et al. 2005; Khan
et al. 2012). Activation of synaptic NMDARs increases association of
CaMKII with spines and the PSD; however, the role and mechanism of this
translocation are still incompletely understood (Khan et al. 2012).
CaMKII is a complex holoenzyme, the structure of which has interesting
consequences for the dynamics of its activation by calcium/CaM (Fig. 3).
Each holoenzyme comprises 12 catalytic subunits held together by their
carboxy-terminal association domains (Bennett et al. 1983; Kolb et al. 1998;
Rosenberg et al. 2005, 2006). Mammalian genomes encode four highly
similar CaMKII subunits: α, β, γ, and δ (Gaertner et al. 2004). They can form
stable homo-oligomers or hetero-oligomers that contain differing numbers of
each isoform; the numbers depend on their relative rates of synthesis. Their
major sequence differences occur in the linker region between the catalytic
and association domains. Only the α and β subunits are highly expressed in
brain; and the α subunit is only expressed in neurons. In forebrain, the α:β
ratio is ∼3:1 (Bennett et al. 1983). Thus, the unusually high levels of CaMKII
in forebrain are primarily a result of the level of expression of the α subunit.
3.2.2 Calcineurin
Calcineurin is a calcium/CaM-activated protein phosphatase found in many
cell types. Injection of inhibitors of calcineurin into postsynaptic neurons in
hippocampal slices first suggested that its activity is required for induction of
LTD (Mulkey et al. 1994). Both brain isoforms of its catalytic subunit
(CNAα and CNAβ) bind calcium-CaM and form a heterodimer with the
regulatory subunit CNB1, which also binds calcium (Kuno et al. 1992). The
requirement of calcineurin for induction of LTD was confirmed in
hippocampal slices from mice with a forebrain-specific deletion of CNB1.
The magnitude of LTD was reduced in these slices and the frequency
threshold for the transition from induction of LTD to induction of LTP was
shifted to a lower value (Zeng et al. 2001). These findings led to the
hypothesis that the relative activation of CaMKII and calcineurin determines
whether LTP or LTD will be induced—a hypothesis that remains to be
proven. An implication of this hypothesis is that the steady-state number of
AMPARs at a synapse and the steady-state size of the actin cytoskeleton are
maintained by a balance of CaMKII and calcineurin activities. More recent
work reveals that the mechanisms controlling the size and strength of the
postsynapse involve the action of several signaling enzymes. For example,
the broader-specificity protein phosphatases PP1 and PP2A also appear to be
necessary for induction of LTP (Mulkey et al. 1993). It is not yet clear how
synaptic activity regulates these two phosphatases. Furthermore, protein
kinases PKA and PKC, which mediate the action of other major second
messenger pathways, can regulate synaptic plasticity when activated by any
of several modulatory neurotransmitters, including acetylcholine, biogenic
amines, and neuropeptides (Abeliovich et al. 1993; Blitzer et al. 1995;
Kennedy et al. 2005).
A possible link between PKA, calcineurin, and PP1 involves a cycle
similar to that found in glycogen metabolism in which a small protein
inhibitor of PP1 (inhibitor 1) is activated by phosphorylation by PKA and
inactivated by dephosphorylation by calcineurin (Huang and Glinsmann
1976; Lisman 1989). However, deletion of inhibitor 1 in the mouse has no
effect on LTP in the Schaffer-collateral pathway or in the medial perforant
pathway that arises in the entorhinal cortex, but reduces LTP in synapses
from the lateral perforant path (Allen et al. 2000). Importantly, the mutation
has no effect on performance of spatial learning tasks. Paralogs of inhibitor 1
that regulate PP1 have been identified in the basal ganglia and cerebellum,
but none has been found in the hippocampus or cortex. The divergent effects
of inhibitor 1 deletion on different synaptic pathways in the hippocampus
show that distinct, heterogeneous molecular mechanisms underlie synaptic
plasticity both in different dendritic subregions and in different neuronal
subtypes. Two other PP1 regulatory proteins, spinophilin and neurabin, can
regulate PP1 activity and its association with the actin cytoskeleton. Deletion
of spinophilin eliminates LTD induced by low-frequency stimulation of the
Schaffer-collateral pathway (Feng et al. 2000). This is consistent with a
requirement for PP1 activity but does not shed light on the relationship of
calcium influx to PP1 activity during induction of LTD.
Calcineurin is a more selective phosphatase than PP1 or PP2A, requiring
upstream motifs, such as LxVP, in its substrates (Grigoriu et al. 2013).
Although there is a small amount of overlap in the target sites for CaMKII
and calcineurin, many of their target sites do not overlap. Therefore, the shift
in phosphorylation status of individual proteins in the spine during and after
an influx of calcium is difficult to predict and is still the subject of study.
3.2.3 Modulatory Calcium-Sensitive Enzymes
The calcium/CaM-stimulated adenylyl cyclase isoform AC1 (Wang and
Storm 2003), the calcium/CaM-activated cyclic nucleotide phosphodiesterase
PDE1 (Sharma et al. 2006), the calcium/CaM-regulated neuronal nitric oxide
synthase isoform (nNOS) (Salerno et al. 2013), members of a family of
calcium-sensitive guanine nucleotide exchange factors (GEFs) called
RasGRF1 and RasGRF2 (Feig 2011), calcium-sensitive PKC isoforms (Lipp
and Reither 2011), and the calcium-dependent protease calpain (Croall and
DeMartino 1991) are all present at low and varying levels in spines and can
modulate the sensitivity of a synapse to induction of synaptic plasticity or the
magnitude and duration of plastic changes. These modulatory mechanisms
are outside our scope here, however.
PSD95 and its three relatives SAP97, SAP102, and PSD93 are centered
∼12–20 nm from the postsynaptic membrane (Valtschanoff and Weinberg
2001; Chen et al. 2008) and link glutamate receptors to the PSD structure and
to proximal signaling enzymes. Each contains three PDZ domains, an SH3
domain, and a carboxy-terminal degenerate guanylate kinase (GuK) domain,
all of which act as protein-docking sites (Kornau et al. 1997; Sheng and Kim
2011). The first two PDZ domains of the PSD95 family can bind to several
synaptic membrane proteins via PDZ-binding motifs in their carboxyl
termini, including the GluN subunits of NMDARs, TARPs, and neuroligin, a
transmembrane adhesion protein. In adult mammals, PSD95 is the most
abundant of the family members and associates preferentially with GluN2A,
whereas SAP102 predominates in synapses during the first few weeks of
development and associates preferentially with GluN2B (Sans et al. 2000).
These two scaffold proteins facilitate the developmental shift from NMDARs
that contain predominantly GluN2B to the adult form that contains a mixture
of GluN2A and GluN2B (Sans et al. 2000; Elias et al. 2006). The transition is
important for synaptic signaling because complexes formed by PSD95 and
SAP102 contain different sets of signaling enzymes.
The PDZ domains can also bind to cytosolic signaling enzymes, including
nNOS and synGAP. The carboxy-terminal SH3 domains of the PSD95 family
have an unusual split structure created by the insertion of the GuK domain
between two α helices (McGee et al. 2001; Tavares et al. 2001). This
structure may help regulate oligomerization at the postsynaptic site. The SH3
domain can also bind directly to an AKAP79/150 scaffold protein that
localizes PKA, PKC, and calcineurin to the PSD (Gold et al. 2011). Finally,
the terminal GuK domain forms a docking site for linker proteins of the
GKAP family (for guanylate kinase-associating protein), providing a critical
bridge between the PSD95 family and the next layer of postsynaptic scaffold
proteins, the SHANK family (Kim et al. 1997; Takeuchi et al. 1997; Naisbitt
et al. 1999).
SHANK proteins act as a “scaffold of scaffolds” within the PSD (Fig. 4)
(see Sheng and Kim 2011). They form an interacting network centered ∼25
nm from the postsynaptic membrane and contain multiple, differentially
spliced protein-interaction domains (Sheng and Kim 2000; Boeckers et al.
2002). GKAP binds to a PDZ domain in SHANK, linking it to the PSD95
scaffold (Fig. 4). A proline-rich domain in SHANK links it to the actin
cytoskeleton via the SH3 domain of cortactin. Cortactin is an F-actin-binding
protein that also enhances binding of actin to the ARP2/3 complex,
facilitating branching of actin filaments.
A second proline-rich domain in SHANK interacts with the Homer family
of scaffold proteins (Fig. 4). Homer proteins form tetramers linked by their
carboxy-terminal coiled-coil domains. Each tetramer contains four identical
amino-terminal EVH domains that bind to proline-rich sites on three classes
of synaptic proteins: SHANK, the cytosolic tails of metabotropic glutamate
receptors (mGluRs), and inositol 1,4,5-trisphosphate (IP3) receptors (IP3Rs).
Thus, the Homer proteins can form a bridge between mGluRs and internal
calcium stores located in vesicles of smooth endoplasmic reticulum that
contain IP3Rs. Not all spines contain such stores, but when they are present
calcium is released into the cytoplasm when IP3 binds to the IP3Rs (p. 95
[Bootman 2012]). Finally, the EVH domains link all of the Homer complexes
directly to SHANK.
The dense protein network formed by these three families of scaffold
proteins is highly dynamic. For example, PSD95 exchanges between
neighboring spines with a median retention time of ∼30–100 min (Gray et al.
2006). Retention time decreases during sensory deprivation (Gray et al. 2006)
and also after synaptic activity (Steiner et al. 2008). The composition and size
of PSDs thus appear to be regulated in an ongoing way by neural activity.
This dynamic scaffold provides the underlying spatial organization for the
signaling events that regulate the strength of the synapse (Fig. 4).
Genome-wide association (GWAS) studies have identified mutations or
copy-number variants in PSD scaffolds as risk factors for autism spectrum
disorders. In humans and in mice, deletion of SHANK scaffold proteins has
been repeatedly linked to autistic behaviors (Durand et al. 2007; Herbert
2011; Peca et al. 2011). Similarly, deletion of PSD95 in mice results in
behaviors associated with autism, and two human single-nucleotide
polymorphisms in the gene encoding PSD95 are associated with
characteristics of William’s syndrome, a genetic disorder that includes a
highly social personality and cognitive difficulties (Feyder et al. 2010).
6 CONCLUDING REMARKS
Synapses in the brain release a number of different neurotransmitters
including GABA, acetylcholine, the biogenic amines serotonin, dopamine,
and norepinephrine, and a wide variety of peptide neurohormones. However,
as far as we now know, it is only excitatory glutamatergic synapses that
display the Hebbian form of regulation discussed here. The sculpting of
excitatory connections in response to input from the environment is the
principal mechanism of memory formation in the brain. As excitatory
connections are altered by the Hebbian mechanism, new neural networks are
formed, and others are weakened or strengthened. All of the other types of
synapses contribute to regulation of Hebbian plasticity and help to determine
the conditions under which specific memories are formed, as well as how
long the memories will last. We know much less about regulation of the size
of the signal and response in these other synaptic types, which are fewer in
number and are dispersed among the more abundant glutamatergic synapses,
making them less accessible to molecular manipulation or measurement.
Another obstacle to our full understanding of synaptic regulation is the
subtle variation in mechanisms of synaptic plasticity in spines of different
excitatory neuronal types and among neurons in different brain regions.
These differences effectively obscure our vision because most experimental
methods either sample blindly from the mixture of synapses in a preparation
or record average changes from a poorly understood mixture of synaptic
types. As we learn which receptors and enzymes play critical roles in
modulating synaptic plasticity, new anatomical techniques such as array
tomography (Micheva et al. 2010) and superresolution confocal microscopy
(Dani et al. 2010) will help to sort out distinct synaptic types.
A final experimental frontier concerns the delicate timing of synaptic
regulation required for healthy brain function. To paraphrase Marc Kirschner
describing regulation of embryonic development, “In the regulation of the
brain, as in the theater, timing is everything. Imagine if, one night, the actors
in a play were to miss every single cue, delivering each line perfectly, but
always too early or too late. The evening would be a disaster. The same is
true in brain function. Starting at the moment when the environment
stimulates sensory endings, neurons in the brain send signals to each other to
coordinate sensory perception, emotional and motor responses, and the
laying down of memories. Not only do the signals have to be correct, they
also must be perfectly timed. Otherwise, disasters like mental illness can
result” (paraphrased words in italics) (see kirschner.med.harvard.edu).
A challenge arises from the fact that the biochemical reactions that initiate
and sculpt changes in spine structure underlying activity-dependent synaptic
plasticity occur in a tiny compartment that contains tens to several hundred
copies of the requisite enzymes and effectors. Some of these are immobilized
by scaffold proteins that hold them in close proximity to the most important
downstream targets. Additional complexity arises from the fact that the
initiating calcium signal is always fluctuating, driven by the stuttering
kinetics of the NMDAR channel and active calcium pumps in the spine
membrane. Thus, time-resolved, high-resolution mass spectroscopy and
engineered biochemical real-time sensors, in concert with modeling methods
such as those afforded by the spatially accurate, stochastic modeling program
MCell (e.g., see Kennedy et al. 2005), will be needed to help resolve rapid,
transient molecular events involved in memory formation from those
underlying homeostatic mechanisms.
REFERENCES
*Reference is in this book.
Abeliovich A, Chen C, Goda Y, Silva AJ, Stevens CF, Tonegawa S. 1993. Modified hippocampal long-
term potentiation in PKCγ-mutant mice. Cell 75: 1253–1262.
Abraham WC, Bear MF. 1996. Metaplasticity: The plasticity of synaptic plasticity. Trends Neurosci
19: 126–130.
Allen PB, Hvalby O, Jensen V, Errington ML, Ramsay M, Chaudhry FA, Bliss TV, Storm-Mathisen J,
Morris RG, Andersen P, et al. 2000. Protein phosphatase-1 regulation in the induction of long-term
potentiation: Heterogeneous molecular mechanisms. J Neurosci 20: 3537–3543.
Ascher P, Nowak L. 1988. The role of divalent cations in the N-methyl-D-aspartate responses of mouse
central neurones in culture. J Physiol 399: 247–266.
Bennett MK, Erondu NE, Kennedy MB. 1983. Purification and characterization of a calmodulin-
dependent protein kinase that is highly concentrated in brain. J Biol Chem 258: 12735–12744.
Blitzer RD, Wong T, Nouranifar R, Iyengar R, Landau EM. 1995. Postsynaptic cAMP pathway gates
early LTP in hippocampal CA1 region. Neuron 15: 1403–1414.
Boeckers TM, Bockmann J, Kreutz MR, Gundelfinger ED. 2002. ProSAP/Shank proteins—a family of
higher order organizing molecules of the postsynaptic density with an emerging role in human
neurological disease. J Neurochem 81: 903–910.
* Bootman MD. 2012. Calcium signaling. Cold Spring Harb Perspect Biol 4: a011171.
Carlisle HJ, Manzerra P, Marcora E, Kennedy MB. 2008. SynGAP regulates steady-state and activity-
dependent phosphorylation of cofilin. J Neurosci 28: 13673–13683.
Carlisle HJ, Luong TN, Medina-Marino A, Schenker LT, Khorosheva EM, Indersmitten T, Gunapala
KM, Steele AD, O’Dell TJ, Patterson PH, et al. 2011. Deletion of densin-180 results in abnormal
behaviors associated with mental illness and reduces mGluR5 and DISC1 in the postsynaptic
density fraction. J Neurosci 31: 16194–16207.
Chao LH, Pellicena P, Deindl S, Barclay LA, Schulman H, Kuriyan J. 2010. Intersubunit capture of
regulatory segments is a component of cooperative CaMKII activation. Nat Struct Mol Biol 17:
264–272.
Chao LH, Stratton MM, Lee IH, Rosenberg OS, Levitz J, Mandell DJ, Kortemme T, Groves JT,
Schulman H, Kuriyan J. 2011. A mechanism for tunable autoinhibition in the structure of a human
Ca2+/calmodulin-dependent kinase II holoenzyme. Cell 146: 732–745.
Chen H-J, Rojas-Soto M, Oguni A, Kennedy MB. 1998. A synaptic Ras-GTPase activating protein
(p135 SynGAP) inhibited by CaM Kinase II. Neuron 20: 895–904.
Chen X, Vinade L, Leapman RD, Petersen JD, Nakagawa T, Phillips TM, Sheng M, Reese TS. 2005.
Mass of the postsynaptic density and enumeration of three key molecules. Proc Natl Acad Sci 102:
11551–11556.
Chen X, Winters C, Azzam R, Li X, Galbraith JA, Leapman RD, Reese TS. 2008. Organization of the
core structure of the postsynaptic density. Proc Natl Acad Sci 105: 4453–4458.
Ch’ng TH, Martin KC. 2011. Synapse-to-nucleus signaling. Curr Opin Neurobiol 21: 345–352.
Croall DE, DeMartino GN. 1991. Calcium-activated neutral protease (calpain) system: Structure,
function, and regulation. Physiol Rev 71: 813–847.
Dani A, Huang B, Bergan J, Dulac C, Zhuang X. 2010. Superresolution imaging of chemical synapses
in the brain. Neuron 68: 843–856.
Ding J-D, Kennedy MB, Weinberg R. 2013. Subcellular organization of CaMKII in rat hippocampal
pyramidal neurons. J Comp Neurol 521: 3570–3583.
Durand CM, Betancur C, Boeckers TM, Bockmann J, Chaste P, Fauchereau F, Nygren G, Rastam M,
Gillberg IC, Anckarsater H, et al. 2007. Mutations in the gene encoding the synaptic scaffolding
protein SHANK3 are associated with autism spectrum disorders. Nat Genet 39: 25–27.
Elias GM, Funke L, Stein V, Grant SG, Bredt DS, Nicoll RA. 2006. Synapse-specific and
developmentally regulated targeting of AMPA receptors by a family of MAGUK scaffolding
proteins. Neuron 52: 307–320.
Erondu NE, Kennedy MB. 1985. Regional distribution of type II Ca2+/calmodulin-dependent protein
kinase in rat brain. J Neurosci 5: 3270–3277.
Feig LA. 2011. Regulation of neuronal function by Ras-GRF exchange factors. Genes Cancer 2: 306–
319.
Feng J, Yan Z, Ferreira A, Tomizawa K, Liauw JA, Zhuo M, Allen PB, Ouimet CC, Greengard P.
2000. Spinophilin regulates the formation and function of dendritic spines. Proc Natl Acad Sci 97:
9287–9292.
Feyder M, Karlsson RM, Mathur P, Lyman M, Bock R, Momenan R, Munasinghe J, Scattoni ML, Ihne
J, Camp M, et al. 2010. Association of mouse Dlg4 (PSD-95) gene deletion and human DLG4 gene
variation with phenotypes relevant to autism spectrum disorders and Williams’ syndrome. Am J
Psychiatry 167: 1508–1517.
Flavell SW, Greenberg ME. 2008. Signaling mechanisms linking neuronal activity to gene expression
and plasticity of the nervous system. Annu Rev Neurosci 31: 563–590.
Foster KA, McLaughlin N, Edbauer D, Phillips M, Bolton A, Constantine-Paton M, Sheng M. 2010.
Distinct roles of NR2A and NR2B cytoplasmic tails in long-term potentiation. J Neurosci 30:
2676–2685.
Franks KM, Sejnowski TJ. 2002. Complexity of calcium signaling in synaptic spines. Bioessays 24:
1130–1144.
Furukawa H, Singh SK, Mancusso R, Gouaux E. 2005. Subunit arrangement and function in NMDA
receptors. Nature 438: 185–192.
Gaertner TR, Kolodziej SJ, Wang D, Kobayashi R, Koomen JM, Stoops JK, Waxham MN. 2004.
Comparative analyses of the three-dimensional structures and enzymatic properties of the α, β, γ
and δ isoforms of Ca2+-calmodulin-dependent protein kinase II. J Biol Chem 279: 12484–12494.
Gardoni F, Schrama LH, Kamal A, Gispen WH, Cattabeni F, Di Luca M. 2001. Hippocampal synaptic
plasticity involves competition between Ca2+/calmodulin-dependent protein kinase II and
postsynaptic density 95 for binding to the NR2A subunit of the NMDA receptor. J Neurosci 21:
1501–1509.
Giese KP, Fedorov NB, Filipkowski RK, Silva AJ. 1998. Autophosphorylation at Thr286 of the α
calcium-calmodulin kinase II in LTP and learning. Science 279: 870–873.
Gold MG, Stengel F, Nygren PJ, Weisbrod CR, Bruce JE, Robinson CV, Barford D, Scott JD. 2011.
Architecture and dynamics of an A-kinase anchoring protein 79 (AKAP79) signaling complex.
Proc Natl Acad Sci 108: 6426–6431.
Gray NW, Weimer RM, Bureau I, Svoboda K. 2006. Rapid redistribution of synaptic PSD-95 in the
neocortex in vivo. PLoS Biol 4: e370.
Grewal SS, York RD, Stork PJ. 1999. Extracellular-signal-regulated kinase signalling in neurons. Curr
Opin Neurobiol 9: 544–553.
Grigoriu S, Bond R, Cossio P, Chen JA, Ly N, Hummer G, Page R, Cyert MS, Peti W. 2013. The
molecular mechanism of substrate engagement and immunosuppressant inhibition of calcineurin.
PLoS Biol 11: e1001492.
Hamdan FF, Gauthier J, Spiegelman D, Noreau A, Yang Y, Pellerin S, Dobrzeniecka S, Cote M,
Perreau-Linck E, Carmant L, et al. 2009. Mutations in SYNGAP1 in autosomal nonsyndromic
mental retardation. N Engl J Med 360: 599–605.
Hamdan FF, Daoud H, Piton A, Gauthier J, Dobrzeniecka S, Krebs MO, Joober R, Lacaille JC, Nadeau
A, Milunsky JM, et al. 2011. De novo SYNGAP1 mutations in nonsyndromic intellectual disability
and autism. Biol Psychiatry 69: 898–901.
Hanson PI, Meyer T, Stryer L, Schulman H. 1994. Dual role of calmodulin in autophosphorylation of
multifunctional CaM kinase may underlie decoding of calcium signals. Neuron 12: 943–956.
Hastie P, Ulbrich MH, Wang HL, Arant RJ, Lau AG, Zhang Z, Isacoff EY, Chen L. 2013. AMPA
receptor/TARP stoichiometry visualized by single-molecule subunit counting. Proc Natl Acad Sci
110: 5163–5168.
* Hemmings BA, Restuccia DF. 2012. PI3K-PKB/Akt pathway. Cold Spring Harb Perspect Biol 4:
a011189.
Herbert MR. 2011. SHANK3, the synapse, and autism. N Engl J Med 365: 173–175.
Herculano-Houzel S. 2009. The human brain in numbers: A linearly scaled-up primate brain. Front
Hum Neurosci 3: 31.
Ho VM, Lee JA, Martin KC. 2011. The cell biology of synaptic plasticity. Science 334: 623–628.
Huang FL, Glinsmann WH. 1976. Separation and characterization of two phosphorylase phosphatase
inhibitors from rabbit skeletal muscle. Eur J Biochem 70: 419–426.
Kennedy MB. 1997. The postsynaptic density at glutamatergic synapses. Trends Neurosci 20: 264–268.
Kennedy MB. 2000. Signal-processing machines at the postsynaptic density. Science 290: 750–754.
Kennedy MB, Bennett MK, Erondu NE. 1983. Biochemical and immunochemical evidence that the
“major postsynaptic density protein” is a subunit of a calmodulin-dependent protein kinase. Proc
Natl Acad Sci 80: 7357–7361.
Kennedy MB, Beale HC, Carlisle HJ, Washburn LR. 2005. Integration of biochemical signalling in
spines. Nat Rev Neurosci 6: 423–434.
Khan S, Reese TS, Rajpoot N, Shabbir A. 2012. Spatiotemporal maps of CaMKII in dendritic spines. J
Comput Neurosci 33: 123–139.
Kim E, Naisbitt S, Hsueh YP, Rao A, Rothschild A, Craig AM, Sheng M. 1997. GKAP, a novel
synaptic protein that interacts with the guanylate kinase-like domain of the PSD-95/SAP90 family
of channel clustering molecules. J Cell Biol 136: 669–678.
Kim JH, Liao D, Lau L-F, Huganir RL. 1998. SynGAP: A synaptic RasGAP that associates with the
PSD-95/SAP90 protein family. Neuron 20: 683–691.
Kim JH, Lee HK, Takamiya K, Huganir RL. 2003. The role of synaptic GTPase-activating protein in
neuronal development and synaptic plasticity. J Neurosci 23: 1119–1124.
Kolb SJ, Hudmon A, Ginsberg TR, Waxham MN. 1998. Identification of domains essential for the
assembly of calcium/calmodulin-dependent protein kinase II holoenzymes. J Biol Chem 273:
31555–31564.
Komiyama NH, Watabe AM, Carlisle HJ, Porter K, Charlesworth P, Monti J, Strathdee DJ, O’Carroll
CM, Martin SJ, Morris RG, et al. 2002. SynGAP regulates ERK/MAPK signaling, synaptic
plasticity, and learning in the complex with postsynaptic density 95 and NMDA receptor. J
Neurosci 22: 9721–9732.
Kornau H-C, Seeburg PH, Kennedy MB. 1997. Interaction of ion channels and receptors with PDZ
domain proteins. Curr Opin Neurobiol 7: 368–373.
Krapivinsky G, Medina I, Krapivinsky L, Gapon S, Clapham DE. 2004. SynGAP-MUPP1-CaMKII
synaptic complexes regulate p38 MAP kinase activity and NMDA receptor-dependent synaptic
AMPA receptor potentiation. Neuron 43: 563–574.
Kuno T, Mukai H, Ito A, Chang CD, Kishima K, Saito N, Tanaka C. 1992. Distinct cellular expression
of calcineurin Aα and Aβ in rat brain. J Neurochem 58: 1643–1651.
Kutsuwada T, Kashiwabuchi N, Mori H, Sakimura K, Kushiya E, Kazuaki A, Meguro H, Masaki H,
Kumanishi T, Arakawa M, et al. 1992. Molecular diversity of the NMDA receptor channel. Nature
358: 36–41.
Lee SJ, Escobedo-Lozoya Y, Szatmari EM, Yasuda R. 2009. Activation of CaMKII in single dendritic
spines during long-term potentiation. Nature 458: 299–304.
Leonard AS, Bayer KU, Merrill MA, Lim IA, Shea MA, Schulman H, Hell JW. 2002. Regulation of
calcium/calmodulin-dependent protein kinase II docking to N-methyl-D-aspartate receptors by
calcium/calmodulin and α-actinin. J Biol Chem 277: 48441–48448.
Lipp P, Reither G. 2011. Protein kinase C: The “masters” of calcium and lipid. Cold Spring Harb
Perspect Biol 3: a004556.
Lisman J. 1989. A mechanism for the Hebb and the anti-Hebb processes underlying learning and
memory. Proc Natl Acad Sci 86: 9574–9578.
Lu W, Shi Y, Jackson AC, Bjorgan K, During MJ, Sprengel R, Seeburg PH, Nicoll RA. 2009. Subunit
composition of synaptic AMPA receptors revealed by a single-cell genetic approach. Neuron 62:
254–268.
Lüscher C, Malenka RC. 2012. NMDA receptor-dependent long-term potentiation and long-term
depression (LTP/LTD). Cold Spring Harb Perspect Biol 4: a005710.
MacDermott AB, Mayer ML, Westbrook GL, Smith SJ, Barker JL. 1986. NMDA-receptor activation
increases cytoplasmic calcium concentration in cultured spinal cord neurones. Nature 321: 519–
522.
Magee JC, Johnston D. 1997. A synaptically controlled, associative signal for Hebbian plasticity in
hippocampal neurons. Science 275: 209–213.
Magee J, Hoffman D, Colbert C, Johnston D. 1998. Electrical and calcium signaling in dendrites of
hippocampal pyramidal neurons. Annu Rev Physiol 60: 327–346.
Makino H, Malinow R. 2009. AMPA receptor incorporation into synapses during LTP: The role of
lateral movement and exocytosis. Neuron 64: 381–390.
Malinow R, Malenka RC. 2002. AMPA receptor trafficking and synaptic plasticity. Annu Rev Neurosci
25: 103–126.
Malinow R, Schulman H, Tsien RW. 1989. Inhibition of post-synaptic PKC or CaMKII blocks
induction but not expression of LTP. Science 245: 862–866.
Mayer ML. 2006. Glutamate receptors at atomic resolution. Nature 440: 456–462.
Mayer ML, Westbrook GL, Guthrie PB. 1984. Voltage-dependent block by Mg2+ of NMDA responses
in spinal cord neurones. Nature 309: 261–263.
McGee AW, Dakoji SR, Olsen O, Bredt DS, Lim WA, Prehoda KE. 2001. Structure of the SH3-
guanylate kinase module from PSD-95 suggests a mechanism for regulated assembly of MAGUK
scaffolding proteins. Mol Cell 8: 1291–1301.
Micheva KD, Busse B, Weiler NC, O’Rourke N, Smith SJ. 2010. Single-synapse analysis of a diverse
synapse population: Proteomic imaging methods and markers. Neuron 68: 639–653.
Miller SG, Kennedy MB. 1986. Regulation of brain type II Ca2+/calmodulin-dependent protein kinase
by autophosphorylation: A Ca2+-triggered molecular switch. Cell 44: 861–870.
Miller SG, Patton BL, Kennedy MB. 1988. Sequences of autophosphorylation sites in neuronal type II
CaM kinase that control Ca2+-independent activity. Neuron 1: 593–604.
Minichiello L. 2009. TrkB signalling pathways in LTP and learning. Nat Rev Neurosci 10: 850–860.
Monyer H, Sprengel R, Schoepfer R, Herb A, Higuchi M, Lomeli H, Burnashev N, Sakmann B,
Seeburg PH. 1992. Heteromeric NMDA receptors: Molecular and functional distinction of
subtypes. Science 256: 1217–1221.
Monyer H, Burnashev N, Laurie DJ, Sakmann B, Seeburg PH. 1994. Developmental and regional
expression in the rat brain and functional properties of four NMDA receptors. Neuron 12: 529–540.
Mulkey RM, Herron CE, Malenka RC. 1993. An essential role for protein phosphatases in hippocampal
long-term depression. Science 261: 1051–1055.
Mulkey RM, Endo S, Shenolikar S, Malenka RC. 1994. Involvement of a calcineurin/inhibitor-1
phosphatase cascade in hippocampal long-term depression. Nature 369: 486–488.
Naisbitt S, Kim E, Tu JC, Xiao B, Sala C, Valtschanoff J, Weinberg RJ, Worley PF, Sheng M. 1999.
Shank, a novel family of postsynaptic density proteins that binds to the NMDA receptor/PSD-
95/GKAP complex and cortactin. Neuron 23: 569–582.
* Newton AC, Bootman MD, Scott JD. 2014. Second messengers. Cold Spring Harb Perspect Biol doi:
10.1101/cshperspect.a005926.
Nowak L, Bregestovski P, Ascher P, Herbet A, Prochiantz A. 1984. Magnesium gates glutamate-
activated channels in mouse central neurones. Nature 307: 462–465.
Oh JS, Chen H-J, Rojas-Soto M, Oguni A, Kennedy MB. 2002. Erratum. Neuron 33: 151.
Oh JS, Manzerra P, Kennedy MB. 2004. Regulation of the neuron-specific Ras GTPase activating
protein, synGAP, by Ca2+/calmodulin-dependent protein kinase II. J Biol Chem 279: 17980–
17988.
Opazo P, Choquet D. 2011. A three-step model for the synaptic recruitment of AMPA receptors. Mol
Cell Neurosci 46: 1–8.
Opazo P, Labrecque S, Tigaret CM, Frouin A, Wiseman PW, De Koninck P, Choquet D. 2010.
CaMKII triggers the diffusional trapping of surface AMPARs through phosphorylation of stargazin.
Neuron 67: 239–252.
Patterson MA, Szatmari EM, Yasuda R. 2010. AMPA receptors are exocytosed in stimulated spines
and adjacent dendrites in a Ras-ERK-dependent manner during long-term potentiation. Proc Natl
Acad Sci 107: 15951–15956.
Patton BL, Miller SG, Kennedy MB. 1990. Activation of type II calcium/calmodulin-dependent protein
kinase by Ca2+/calmodulin is inhibited by autophosphorylation of threonine within the calmodulin-
binding domain. J Biol Chem 265: 11204–11212.
Peca J, Feliciano C, Ting JT, Wang W, Wells MF, Venkatraman TN, Lascola CD, Fu Z, Feng G. 2011.
Shank3 mutant mice display autistic-like behaviours and striatal dysfunction. Nature 472: 437–442.
Penzes P, Woolfrey KM, Srivastava DP. 2011. Epac2-mediated dendritic spine remodeling:
Implications for disease. Mol Cell Neurosci 46: 368–380.
Peters A, Kaiserman-Abramof IR. 1970. The small pyramidal neurons of the cerebral cortex. The
perikaryon, dendrites and spines. Am J Anat 127: 321–356.
Petrini EM, Lu J, Cognet L, Lounis B, Ehlers MD, Choquet D. 2009. Endocytic trafficking and
recycling maintain a pool of mobile surface AMPA receptors required for synaptic potentiation.
Neuron 63: 92–105.
Rellos P, Pike ACW, Niesen FH, Salah E, Lee WH, von Delft F, Knapp S. 2010. Structure of the
CaMKIIδ/calmodulin complex reveals the molecular mechanism of CaMKII kinase activation.
PLoS Biol 8: e1000426.
Rosenberg OS, Deindl S, Sung RJ, Nairn AC, Kuriyan J. 2005. Structure of the autoinhibited kinase
domain of CaMKII and SAXS analysis of the holoenzyme. Cell 123: 849–860.
Rosenberg OS, Deindl S, Comolli LR, Hoelz A, Downing KH, Nairn AC, Kuriyan J. 2006.
Oligomerization states of the association domain and the holoenyzme of Ca/CaM kinase II. FEBS J
273: 682–694.
Rumbaugh G, Adams JP, Kim JH, Huganir RL. 2006. SynGAP regulates synaptic strength and
mitogen-activated protein kinases in cultured neurons. Proc Natl Acad Sci 103: 4344–4351.
Salerno JC, Ray K, Poulos T, Li H, Ghosh DK. 2013. Calmodulin activates neuronal nitric oxide
synthase by enabling transitions between conformational states. FEBS Lett 587: 44–47.
Salter MW, Kalia LV. 2004. Src kinases: A hub for NMDA receptor regulation. Nat Rev Neurosci 5:
317–328.
Sans N, Petralia RS, Wang YX, Blahos J, Hell JW, Wenthold RJ. 2000. A developmental change in
NMDA receptor-associated proteins at hippocampal synapses. J Neurosci 20: 1260–1271.
* Sassone-Corsi P. 2012. The cyclic AMP pathway. Cold Spring Harb Perspect Biol 4: a011148.
Schworer CM, Colbran RJ, Keefer JR, Soderling TR. 1988. Ca2+/calmodulin-dependent protein kinase
II. Identification of a regulatory autophosphorylation site adjacent to the inhibitory and calmodulin-
binding domains. J Biol Chem 263: 13486–13489.
Seeburg PH. 1993. The molecular biology of mammalian glutamate receptor channels. Trends Neurosci
16: 359–365.
Sharma RK, Das SB, Lakshmikuttyamma A, Selvakumar P, Shrivastav A. 2006. Regulation of
calmodulin-stimulated cyclic nucleotide phosphodiesterase (PDE1): Review. Int J Mol Med 18: 95–
105.
Shen K, Meyer T. 1999. Dynamic control of CaMKII translocation and localization in hippocampal
neurons by NMDA receptor stimulation. Science 284: 162–166.
Shen K, Teruel MN, Subramanian K, Meyer T. 1998. CaMKIIβ functions as an F-actin targeting
module that localizes CaMKIIα/β heterooligomers to dendritic spines. Neuron 21: 593–606.
Sheng M, Kim E. 2000. The Shank family of scaffold proteins. J Cell Sci 113: 1851–1856.
Sheng M, Kim E. 2011. The postsynaptic organization of synapses. Cold Spring Harb Perspect Biol 3:
a005678.
Shi S, Hayashi Y, Esteban JA, Malinow R. 2001. Subunit-specific rules governing AMPA receptor
trafficking to synapses in hippocampal pyramidal neurons. Cell 105: 331–343.
Shi G-X, Han J, Andres DA. 2005. Rin GTPase couples nerve growth factor signaling to p38 and b-
Raf/ERK pathways to promote neuronal differentiation. J Biol Chem 280: 37599–37609.
Shields SM, Ingebritsen TS, Kelly PT. 1985. Identification of protein phosphatase-1 in synaptic
junctions: Dephosphorylaton of endogenous calmodulin-dependent kinase II and synapse-enriched
phosphoproteins. J Neurosci 5: 3414–3422.
Sik A, Hajos N, Gulacsi A, Mody I, Freund TF. 1998. The absence of a major Ca2+ signaling pathway
in GABAergic neurons of the hippocampus. Proc Natl Acad Sci 95: 3245–3250.
Silva AJ, Paylor R, Wehner JM, Tonegawa S. 1992a. Impaired spatial learning in α-calcium-
calmodulin kinase II mutant mice. Science 257: 206–211.
Silva AJ, Stevens CF, Tonegawa S, Wang Y. 1992b. Deficient hippocampal long-term potentiation in
α-calcium-calmodulin kinase II mutant mice. Science 257: 201–206.
Sjostrom PJ, Nelson SB. 2002. Spike timing, calcium signals and synaptic plasticity. Curr Opin
Neurobiol 12: 305–314.
Spruston N, Schiller Y, Stuart G, Sakmann B. 1995. Activity-dependent action potential invasion and
calcium influx into hippocampal CA1 dendrites. Science 268: 297–300.
Steiner P, Higley MJ, Xu W, Czervionke BL, Malenka RC, Sabatini BL. 2008. Destabilization of the
postsynaptic density by PSD-95 serine 73 phosphorylation inhibits spine growth and synaptic
plasticity. Neuron 60: 788–802.
Stork PJ. 2003. Does Rap1 deserve a bad Rap? Trends Biochem Sci 28: 267–275.
Stuart G, Spruston N, Sakmann B, Häusser M. 1997. Action potential initiation and backpropagation in
neurons of the mammalian CNS. Trends Neurosci 20: 125–131.
Takeuchi M, Hata Y, Hirao K, Toyoda A, Irie M, Takai Y. 1997. SAPAPs. A family of PSD-
95/SAP90-associated proteins localized at postsynaptic density. J Biol Chem 272: 11943–11951.
Tavares GA, Panepucci EH, Brunger AT. 2001. Structural characterization of the intramolecular
interaction between the SH3 and guanylate kinase domains of PSD-95. Mol Cell 8: 1313–1325.
Tolias KF, Bikoff JB, Burette A, Paradis S, Harrar D, Tavazoie S, Weinberg RJ, Greenberg ME. 2005.
The Rac1-GEF Tiam1 couples the NMDA receptor to the activity-dependent development of
dendritic arbors and spines. Neuron 45: 525–538.
Tomita S, Stein V, Stocker TJ, Nicoll RA, Bredt DS. 2005. Bidirectional synaptic plasticity regulated
by phosphorylation of stargazin-like TARPs. Neuron 45: 269–277.
Valtschanoff JG, Weinberg RJ. 2001. Laminar organization of the NMDA receptor complex within the
postsynaptic density. J Neurosci 21: 1211–1217.
Vazquez LE, Chen HJ, Sokolova I, Knuesel I, Kennedy MB. 2004. SynGAP regulates spine formation.
J Neurosci 24: 8862–8872.
Walikonis RS, Oguni A, Khorosheva EM, Jeng C-J, Asuncion FJ, Kennedy MB. 2001. Densin-180
forms a ternary complex with the α-subunit of CaMKII and α-actinin. J Neurosci 21: 423–433.
Wang H, Storm DR. 2003. Calmodulin-regulated adenylyl cyclases: Cross-talk and plasticity in the
central nervous system. Mol Pharmacol 63: 463–468.
Yang K, Trepanier C, Sidhu B, Xie YF, Li H, Lei G, Salter MW, Orser BA, Nakazawa T, Yamamoto
T, et al. 2012. Metaplasticity gated through differential regulation of GluN2A versus GluN2B
receptors by Src family kinases. EMBO J 31: 805–816.
York RD, Yao H, Dillon T, Ellig CL, Eckert SP, McCleskey EW, Stork PJ. 1998. Rap1 mediates
sustained MAP kinase activation induced by nerve growth factor. Nature 392: 622–626.
Yoshii A, Constantine-Paton M. 2010. Postsynaptic BDNF-TrkB signaling in synapse maturation,
plasticity, and disease. Dev Neurobiol 70: 304–322.
Zeng H, Chattarji S, Barbarosie M, Rondi-Reig L, Philpot BD, Miyakawa T, Bear MF, Tonegawa S.
2001. Forebrain-specific calcineurin knockout selectively impairs bidirectional synaptic plasticity
and working/episodic-like memory. Cell 107: 617–629.
Zhu JJ, Qin Y, Zhao M, Van Aelst L, Malinow R. 2002. Ras and Rap control AMPA receptor
trafficking during synaptic plasticity. Cell 110: 443–455.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a016824
CHAPTER 13
SUMMARY
Outline
1 Introduction
2 Skeletal muscle contraction
3 Skeletal muscle fiber types and exercise
4 Malignant hyperthermia in skeletal muscle
5 Cardiac muscle contraction
6 Exercise hypertrophy in cardiac muscle
7 Pathophysiological cardiac hypertrophy
8 Heart failure
9 Smooth muscle types
10 The contractile process in smooth muscle
11 Calcium sensitization
12 Vascular smooth muscle in disease
13 Concluding remarks
References
1 INTRODUCTION
Muscle can be subdivided into two general categories: striated muscle, which
includes skeletal and cardiac muscles; and nonstriated muscle, which
includes smooth muscle such as vascular, respiratory, uterine, and
gastrointestinal muscles. In all muscle types, the contractile apparatus
consists of two main proteins: actin and myosin. Striated muscle is so called
because the regular arrangement of alternating actomyosin fibers gives it a
striped appearance. This arrangement allows coordinated contraction of the
whole muscle in response to neuronal stimulation through a voltage- and
calcium-dependent process known as excitation–contraction coupling. The
coupling enables the rapid and coordinated contraction required of skeletal
muscles and the heart. Smooth muscle does not contain regular striations or
undergo the same type of excitation–contraction coupling. Instead, it
typically uses second messenger signaling to open intracellular channels that
release the calcium ions that control the contractile apparatus. These
processes, in contrast to excitation–contraction coupling, are slow and thus
suitable for the slower and more sustained contractions required of smooth
muscle. The actomyosin contractile apparatus is both calcium- and
phosphorylation-dependent, and restoration of basal calcium levels or its
phosphorylation status returns an actively contracting muscle to a
noncontractile state. Muscle-specific signals modulate these processes,
depending on the type of muscle, its function, and the amount of force
required.
In all muscle cells, contraction thus depends on an increase in cytosolic
calcium concentration (Fig. 1). Calcium has an extracellular concentration of
2–4 mM and a resting cytosolic concentration of ∼100 nM. It is also stored
inside cells within the sarcoplasmic (SR, referring to skeletal and cardiac
muscle) and endoplasmic reticulum (ER, referring to smooth muscle) at a
concentration of ∼0.4 mM (p. 95 [Bootman 2012]). In striated muscle, the
increase in calcium levels is due to its release from the SR stores via
ryanodine receptor (RyRs). Neurotransmitters such as acetylcholine bind to
receptors on the muscle surface and elicit a depolarization by causing
sodium/calcium ions to enter through associated channels. This shifts the
resting membrane potential to a more positive value, which in turn activates
voltage-gated channels, resulting in an action potential (the “excitation” part).
The action potential stimulates L-type calcium channels (also known as
dihydropyridine receptors). In skeletal muscle, these are mechanically
coupled to the SR RyRs and open them directly. In cardiac muscle, calcium
influx through the L-type channels opens RyRs via calcium-induced calcium
release (CICR) (p. 95 [Bootman 2012]). The RyR is a large tetrameric six-
transmembrane-span calcium-release channel. Of the three RyR subtypes,
RyR1 is predominantly found in skeletal muscle (see review by Klein et al.
1996), and RyR2 is predominantly found in cardiac muscle (Cheng et al.
1993).
Figure 1. Overview of muscle contraction signals in striated (A) and smooth (B) muscle.
Figure 3. Skeletal muscle contraction and changes with exercise. (A) Neurotransmitter (acetylcholine,
ACh) released from nerve endings binds to receptors (AChRs) on the muscle surface. The ensuing
depolarization causes sodium channels to open, which elicits an action potential that propagates along
the cell. The action potential invades T-tubules and causes the L-type calcium channels to open, which
in turn causes ryanodine receptors (RyRs) in the SR to open and release calcium, which stimulates
contraction. Calcium is pumped back into the SR by (SR/ER calcium ATPase SERCA) pumps. The
decreasing cytosolic calcium levels cause calcium to disassociate from troponin C and, consequently,
tropomyosin reverts to a conformation that covers the myosin-binding sites. (B) Signaling in exercised
skeletal muscle. Both calcium and calcium-independent signals stimulate the transcriptional coactivator
PGC1α. This activates a number of transcription factors that regulate genes associated with
mitochondrial biogenesis, glucose, and lipid homeostasis.
The action potential runs along the top of the muscle and invades the T-
tubules (specialized invaginations of the membrane containing numerous ion
channels). The opening of voltage-gated sodium channels activates L-type
voltage-gated calcium channels lining the T-tubule. A conformational change
in these enables release of calcium on the closely apposed SR via activation
of RyR1. Calcium then binds to troponin as described above, initiating the
contraction process. Calcium-bound CaM also activates MLCK, whose
phosphorylation of the MLC changes cross-bridge properties. This modulates
the troponin-dependent contraction, although there is no effect on the ATPase
activity of MLC. MLC phosphorylation instead enhances force development
at submaximal saturating calcium concentrations (see below). The phosphate
group is subsequently removed by protein phosphatase 1 (PP1).
Figure 4. Cardiac muscle contraction and changes with exercise. (A) Cardiac muscle contraction can
occur as a consequence of calcium entry through L-type calcium channels, which activate ryanodine
receptor (RyR) channels in the SR. Alternatively, β-adrenergic receptors on the cell membrane lead to
activation of adenylyl cyclase (AC), which stimulates PKA. This can promote contraction by
phosphorylating RyR and L-type calcium channels or relaxation by phosphorylating the SERCA pump
inhibitor phospholamban. (B) Changes with exercise lead to an activation of the PI3K/Akt pathway,
and a down-regulation of NFAT and calcinurin.
Figure 5. Smooth muscle contraction. Calcium released by L-type calcium channels or IP3Rs
downstream from Gq-coupled cell-surface receptors causes smooth muscle contraction. It binds to
calmodulin (CaM) and the resulting complex stimulates myosin light-chain (MLC) kinase (MLCK).
This phosphorylates MLC to promote contraction. A RhoA/ROCK pathway and a diacylglycerol
(DAG) pathway contribute to calcium sensitization by altering the phosphorylation status of myosin
light-chain phosphatase (MLCP). Relaxation is mediated through the cGMP/PKG pathway downstream
from nitric oxide (NO) and agonists such as atrial natriuretic peptide (ANP).
11 CALCIUM SENSITIZATION
Calcium sensitization is an essentially calcium-independent process that
enables the amount of constriction in smooth muscle to be tuned by an
alteration in the sensitivity of MLC to calcium (Fig. 5). This process enables
the muscle to sustain a contraction once the initial calcium transient has
dissipated. There are two mechanisms for calcium sensitization: a DAG-PLC-
PKC pathway and a RhoA pathway (Lincoln 2007).
Diacylglycerol (DAG) is produced by PLCβ downstream from certain
GPCRs and activates the conventional and novel protein kinase C (cPKC and
nPKC), but not atypical PKC (aPKC) (Steinberg 2008). PKC has a variety of
downstream targets, such as MLCK and C-kinase potentiated protein
phosphatase 1 inhibitor, molecular mass 17 kDa (CPI-17), both of which
enhance constriction. CPI-17 is a smooth-muscle-specific inhibitor of MLCP
that binds to its catalytic subunit and inhibits phosphatase activity, allowing
contraction to persist.
Several agonists, including angiotensin II, norepinephrine, and endothelin,
activate the small G protein RhoA. RhoA in turn activates Rho kinase
(ROCK), which can mediate calcium sensitization through two main
pathways. First, ROCK stimulates phosphorylation of MYPT1 (Feng et al.
1999). This can be direct, at T695 or T853, with a preference for T853.
Alternatively, it can phosphorylate another kinase, zipper-interacting protein
kinase (ZIPK, also known as DAPK3), which phosphorylates MYPT1
primarily at T695 (Kiss et al. 2002). ZIPK also phosphorylates MLC at
T18/S19. Phosphorylation of MYPT1 interferes with binding of MLCP to
MLC, and thus is believed to decrease phosphatase activity. ROCK can also
phosphorylate CPI-17 (MacDonald et al. 2001).
The preference for the MYPT1 or CPI-17 pathway depends on the type of
smooth muscle. Whereas MYPT1 is ubiquitously expressed in smooth
muscle, CPI-17 is differentially expressed. Moreover, RhoA and associated
proteins are expressed at lower levels in phasic smooth muscle compared
with tonic smooth muscle (Patel and Rattan 2006). Note that PKC can also
phosphorylate CPI-17 to prevent MLCP activity. Within resistance arteries,
an increase in vascular pressure also activates the RhoA pathway; however,
the signaling intermediates linking the change in vascular pressure and the
activation of RhoA remain unknown (Cole and Welsh 2011).
12 VASCULAR SMOOTH MUSCLE IN
DISEASE
Smooth muscle cells are remarkably plastic, altering their phenotype in
response to conditions such as vascular injury, altered blood flow conditions,
or disease states. The changes in phenotype that can occur include cell
proliferation, apoptosis, and cell migration and are induced by many factors,
including cytokines and growth factors, mechanical forces, neuronal stimuli,
and genetic factors. Here we limit our discussion to hypertension.
In hypertension, there is often a change in the sympathetic nervous system
and the renin–angiotensin system that leads to increased blood pressure.
Angiotensinogen is converted to angiotensin I by renin, which in turn is
converted to angiotensin II by angiotensin-converting enzyme (ACE).
Increased circulating angiotensin II acts on the angiotensin receptors (AT1
and AT2), which, when activated, cause increased peripheral resistance. The
consequence for smooth muscle cells is they become hypercontractile.
Treatments include ACE inhibitors (which inhibit the conversion of
angiotensin I to angiotensin II), α1-adrenergic antagonists (which block the
AT1 and AT2 GPCRs), and calcium channel blockers (such as
dihydropyridines, which inhibit the voltage-gated calcium channels). All of
these treatments aim to reduce the contractility of smooth muscle. Interfering
with downstream targets such as RhoA signaling in hypertensive animals has
also been shown to be effective (Uehata et al. 1997; Seko et al. 2003; Moriki
et al. 2004).
The sustained contractile state of vascular smooth muscle is associated
with the activation of calcium-dependent transcription factors. These include
SRF, FOS, NFAT, and CREB. SRF, which is activated by the RhoA
pathway, promotes the expression of genes encoding components of the
contractile apparatus. Calcium-stimulated CaMKII activates and causes the
translocation of CaMKIV to the nucleus, where it can activate CREB, which
promotes transcription of components of the contractile apparatus and other
targets. However, CaMKII can also activate a phosphatase that
dephosphorylates and thus inactivates CREB (Matchkov et al. 2012). NFAT
is activated on dephosphorylation by calcium-activated calcineurin, which
induces genes associated with proliferation and migration.
NO produced by eNOS in endothelial cells protects against the changes
observed in hypertension: the cGMP pathway inhibits DNA synthesis,
mitogenesis, and cell proliferation (Forstermann and Sessa 2012). However,
endothelial dysfunction is a hallmark of vascular disease, including
hypertension. In many types of vascular diseases, eNOS is up-regulated but
owing to reduced oxygen availability it is converted to a dysfunctional
enzyme that produces superoxides, which contribute to vascular oxidative
stress (Forstermann and Sessa 2012).
In some disease states, smooth muscle cells adopt a noncontractile
phenotype. Although these cells still have signaling machinery that increases
intracellular calcium levels, they have significantly reduced calcium influx
through voltage-gated calcium channels. Thus, there is a shift to intracellular-
store-operated calcium release, similar to the changes observed in cardiac
hypertrophy. Concomitant with decreases in the levels of SERCA, RyR2,
PMCA1, and the sodium/calcium exchanger, the levels of STIM, ORAI
(proteins associated with refilling of intracellular calcium stores; see p. 95
[Bootman 2012]), SERCA2B, and IP3R increase and there is a change in RyR
receptor subtypes from RyR2 to RyR3 (Lipskaia and Lompre 2004; Berra-
Romani et al. 2008; Baryshnikov et al. 2009; Matchkov et al. 2012). These
changes collectively reflect a less contractile phenotype.
13 CONCLUDING REMARKS
Signal transduction is essential for the function of contractile cells. The
stimulatory signal results in an increase in cytosolic calcium levels, which
activates muscle contraction. We now know the main contributors to the
various types of muscle contraction, and have a better appreciation of the
changes that occur to the contractile apparatus under exercise and
pathophysiological conditions. For example, the identification of PGC1α as a
master regulator of transcription factors up-regulated in both exercised and
pathological striated muscle provides new avenues to modulate muscles in a
therapeutic setting. It is also apparent that many signaling proteins in both
smooth and striated muscles are activated by changes in cytosolic calcium
levels, and these signaling pathways often lead to alterations in gene
expression. Because we now have a better appreciation of the changes that
occur to the contractile apparatus under pathophysiological conditions, this
knowledge can be harnessed to allow us to treat disease strategically.
ACKNOWLEDGMENTS
Research in the Ehrlich Laboratory is supported by National Institutes of
Health funds. I.Y.K. is an American Heart Association postdoctoral fellow.
REFERENCES
*Reference is in this book.
Andersson DC, Marks AR. 2010. Fixing ryanodine receptor Ca leak—A novel therapeutic strategy for
contractile failure in heart and skeletal muscle. Drug Discov Today 7: e151–e157.
Balakumar P, Jagadeesh G. 2010. Multifarious molecular signaling cascades of cardiac hypertrophy:
Can the muddy waters be cleared? Pharmacol Res 62: 365–383.
Baryshnikov SG, Pulina MV, Zulian A, Linde CI, Golovina VA. 2009. Orai1, a critical component of
store-operated Ca2+ entry, is functionally associated with Na+/Ca2+ exchanger and plasma
membrane Ca2+ pump in proliferating human arterial myocytes. Am J Physiol Cell Physiol 297:
C1103–C1112.
Berra-Romani R, Mazzocco-Spezzia A, Pulina MV, Golovina VA. 2008. Ca2+ handling is altered
when arterial myocytes progress from a contractile to a proliferative phenotype in culture. Am J
Physiol Cell Physiol 295: C779–C790.
Bezprozvanny I, Watras J, Ehrlich BE. 1991. Bell-shaped calcium-response curves of Ins(1,4,5)P3- and
calcium-gated channels from endoplasmic reticulum of cerebellum. Nature 351: 751–754.
* Bootman MD. 2012. Calcium signaling. Cold Spring Harb Perpsect Biol 4: a011171.
Bozzo C, Spolaore B, Toniolo L, Stevens L, Bastide B, Cieniewski-Bernard C, Fontana A, Mounier Y,
Reggiani C. 2005. Nerve influence on myosin light chain phosphorylation in slow and fast skeletal
muscles. FEBS J 272: 5771–5785.
Campbell KP, MacLennan DH, Jorgensen AO, Mintzer MC. 1983. Purification and characterization of
calsequestrin from canine cardiac sarcoplasmic reticulum and identification of the 53,000 dalton
glycoprotein. J Biol Chem 258: 1197–1204.
Cheng H, Lederer WJ, Cannell MB. 1993. Calcium sparks: Elementary events underlying excitation-
contraction coupling in heart muscle. Science 262: 740–744.
Chien KR. 1999. Stress pathways and heart failure. Cell 98: 555–558.
Cole WC, Welsh DG. 2011. Role of myosin light chain kinase and myosin light chain phosphatase in
the resistance arterial myogenic response to intravascular pressure. Arch Biochem Biophys 510:
160–173.
Dainese M, Quarta M, Lyfenko AD, Paolini C, Canato M, Reggiani C, Dirksen RT, Protasi F. 2009.
Anesthetic- and heat-induced sudden death in calsequestrin-1-knockout mice. FASEB J 23: 1710–
1720.
D’Angelo DD, Sakata Y, Lorenz JN, Boivin GP, Walsh RA, Liggett SB, Dorn GW 2nd. 1997.
Transgenic Gαq overexpression induces cardiac contractile failure in mice. Proc Natl Acad Sci 94:
8121–8126.
Feng J, Ito M, Ichikawa K, Isaka N, Nishikawa M, Hartshorne DJ, Nakano T. 1999. Inhibitory
phosphorylation site for Rho-associated kinase on smooth muscle myosin phosphatase. J Biol Chem
274: 37385–37390.
Finch EA, Turner TJ, Goldin SM. 1991. Calcium as a coagonist of inositol 1,4,5-trisphosphate-induced
calcium release. Science 252: 443–446.
Forstermann U, Sessa WC. 2012. Nitric oxide synthases: Regulation and function. Eur Heart J 33:
829–837, 837a–837d.
Handschin C, Spiegelman BM. 2006. Peroxisome proliferator-activated receptor γ coactivator 1
coactivators, energy homeostasis, and metabolism. Endocr Rev 27: 728–735.
* Hardie DG. 2012. Organismal carbohydrate and lipid homeostasis. Cold Spring Harb Perspect Biol
4: a006031.
Harzheim D, Talasila A, Movassagh M, Foo RS, Figg N, Bootman MD, Roderick HL. 2010. Elevated
InsP3R expression underlies enhanced calcium fluxes and spontaneous extra-systolic calcium
release events in hypertrophic cardiac myocytes. Channels (Austin) 4: 67–71.
Hasenfuss G, Reinecke H, Studer R, Meyer M, Pieske B, Holtz J, Holubarsch C, Posival H, Just H,
Drexler H. 1994. Relation between myocardial function and expression of sarcoplasmic reticulum
Ca2+-ATPase in failing and nonfailing human myocardium. Circ Res 75: 434–442.
He WQ, Peng YJ, Zhang WC, Lv N, Tang J, Chen C, Zhang CH, Gao S, Chen HQ, Zhi G, et al. 2008.
Myosin light chain kinase is central to smooth muscle contraction and required for gastrointestinal
motility in mice. Gastroenterology 135: 610–620.
* Hemmings BA, Restuccia DF. 2012. The PI3K-PKB/Akt pathway. Cold Spring Harb Perspect Biol
4: a011189.
Ibrahim M, Gorelik J, Yacoub MH, Terracciano CM. 2011. The structure and function of cardiac t-
tubules in health and disease. Proc Biol Sci 278: 2714–2723.
Kiss E, Muranyi A, Csortos C, Gergely P, Ito M, Hartshorne DJ, Erdodi F. 2002. Integrin-linked kinase
phosphorylates the myosin phosphatase target subunit at the inhibitory site in platelet cytoskeleton.
Biochem J 365: 79–87.
Klein MG, Cheng H, Santana LF, Jiang YH, Lederer WJ, Schneider MF. 1996. Two mechanisms of
quantized calcium release in skeletal muscle. Nature 379: 455–458.
Kuo IY, Wolfle SE, Hill CE. 2011. T-type calcium channels and vascular function: The new kid on the
block? J Physiol 589: 783–795.
Lanner JT, Georgiou DK, Joshi AD, Hamilton SL. 2010. Ryanodine receptors: Structure, expression,
molecular details, and function in calcium release. Cold Spring Harb Perspect Biol 2: a003996.
Lin J, Wu H, Tarr PT, Zhang CY, Wu Z, Boss O, Michael LF, Puigserver P, Isotani E, Olson EN, et al.
2002. Transcriptional co-activator PGC-1α drives the formation of slow-twitch muscle fibres.
Nature 418: 797–801.
Lincoln TM. 2007. Myosin phosphatase regulatory pathways: Different functions or redundant
functions? Circ Res 100: 10–12.
Lipskaia L, Lompre AM. 2004. Alteration in temporal kinetics of Ca2+ signaling and control of growth
and proliferation. Biol Cell 96: 55–68.
Liu Y, Shen T, Randall WR, Schneider MF. 2005. Signaling pathways in activity-dependent fiber type
plasticity in adult skeletal muscle. J Muscle Res Cell Motil 26: 13–21.
MacDonald JA, Eto M, Borman MA, Brautigan DL, Haystead TA. 2001. Dual Ser and Thr
phosphorylation of CPI-17, an inhibitor of myosin phosphatase, by MYPT-associated kinase. FEBS
Lett 493: 91–94.
Matchkov VV, Kudryavtseva O, Aalkjaer C. 2012. Intracellular Ca2+ signalling and phenotype of
vascular smooth muscle cells. Basic Clin Pharmacol Toxicol 110: 42–48.
Matsui T, Nagoshi T, Rosenzweig A. 2003. Akt and PI 3-kinase signaling in cardiomyocyte
hypertrophy and survival. Cell Cycle 2: 220–223.
Meyer M, Schillinger W, Pieske B, Holubarsch C, Heilmann C, Posival H, Kuwajima G, Mikoshiba K,
Just H, Hasenfuss G, et al. 1995. Alterations of sarcoplasmic reticulum proteins in failing human
dilated cardiomyopathy. Circulation 92: 778–784.
Mishra S, Ling H, Grimm M, Zhang T, Bers DM, Brown JH. 2010. Cardiac hypertrophy and heart
failure development through Gq and CaM kinase II signaling. J Cardiovasc Pharmacol 56: 598–
603.
Miyata S, Minobe W, Bristow MR, Leinwand LA. 2000. Myosin heavy chain isoform expression in the
failing and nonfailing human heart. Circ Res 86: 386–390.
Moriki N, Ito M, Seko T, Kureishi Y, Okamoto R, Nakakuki T, Kongo M, Isaka N, Kaibuchi K,
Nakano T. 2004. RhoA activation in vascular smooth muscle cells from stroke-prone spontaneously
hypertensive rats. Hypertens Res 27: 263–270.
* Morrison DK. 2012. MAP kinase pathways. Cold Spring Harb Perspect Biol 4: a0011254.
Nikolaev VO, Moshkov A, Lyon AR, Miragoli M, Novak P, Paur H, Lohse MJ, Korchev YE, Harding
SE, Gorelik J. 2010. β2-Adrenergic receptor redistribution in heart failure changes cAMP
compartmentation. Science 327: 1653–1657.
Nishikimi T, Kuwahara K, Nakao K. 2011. Current biochemistry, molecular biology, and clinical
relevance of natriuretic peptides. J Cardiol 57: 131–140.
Olesen J, Kiilerich K, Pilegaard H. 2010. PGC-1α-mediated adaptations in skeletal muscle. Pflugers
Archiv Eur J Phys 460: 153–162.
Oliveira RS, Ferreira JC, Gomes ER, Paixao NA, Rolim NP, Medeiros A, Guatimosim S, Brum PC.
2009. Cardiac anti-remodelling effect of aerobic training is associated with a reduction in the
calcineurin/NFAT signalling pathway in heart failure mice. J Physiol 587: 3899–3910.
Patel CA, Rattan S. 2006. Spontaneously tonic smooth muscle has characteristically higher levels of
RhoA/ROK compared with the phasic smooth muscle. Am J Physiol Gastrointest Liver Physiol
291: G830–G837.
Postma AV, Denjoy I, Hoorntje TM, Lupoglazoff JM, Da Costa A, Sebillon P, Mannens MM, Wilde
AA, Guicheney P. 2002. Absence of calsequestrin 2 causes severe forms of catecholaminergic
polymorphic ventricular tachycardia. Circ Res 91: e21–e26.
Pritchard TJ, Kranias EG. 2009. Junctin and the histidine-rich Ca2+ binding protein: Potential roles in
heart failure and arrhythmogenesis. J Physiol 587: 3125–3133.
Robinson R, Carpenter D, Shaw MA, Halsall J, Hopkins P. 2006. Mutations in RYR1 in malignant
hyperthermia and central core disease. Hum Mutat 27: 977–989.
Ronnebaum SM, Patterson C. 2010. The FoxO family in cardiac function and dysfunction. Ann Rev
Physiol 72: 81–94.
Sanbe A, Fewell JG, Gulick J, Osinska H, Lorenz J, Hall DG, Murray LA, Kimball TR, Witt SA,
Robbins J. 1999. Abnormal cardiac structure and function in mice expressing nonphosphorylatable
cardiac regulatory myosin light chain 2. J Biol Chem 274: 21085–21094.
Schiaffino S, Reggiani C. 2011. Fiber types in mammalian skeletal muscles. Physiol Rev 91: 1447–
1531.
Schlecker C, Boehmerle W, Jeromin A, DeGray B, Varshney A, Sharma Y, Szigeti-Buck K, Ehrlich
BE. 2006. Neuronal calcium sensor-1 enhancement of InsP3 receptor activity is inhibited by
therapeutic levels of lithium. J Clin Invest 116: 1668–1674.
Seko T, Ito M, Kureishi Y, Okamoto R, Moriki N, Onishi K, Isaka N, Hartshorne DJ, Nakano T. 2003.
Activation of RhoA and inhibition of myosin phosphatase as important components in hypertension
in vascular smooth muscle. Circ Res 92: 411–418.
Steinberg SF. 2008. Structural basis of protein kinase C isoform function. Physiol Rev 88: 1341–1378.
Uehata M, Ishizaki T, Satoh H, Ono T, Kawahara T, Morishita T, Tamakawa H, Yamagami K, Inui J,
Maekawa M, et al. 1997. Calcium sensitization of smooth muscle mediated by a ρ-associated
protein kinase in hypertension. Nature 389: 990–994.
Ventura-Clapier R, Mettauer B, Bigard X. 2007. Beneficial effects of endurance training on cardiac and
skeletal muscle energy metabolism in heart failure. Cardiovasc Res 73: 10–18.
Watson PA, Reusch JE, McCune SA, Leinwand LA, Luckey SW, Konhilas JP, Brown DA, Chicco AJ,
Sparagna GC, Long CS, et al. 2007. Restoration of CREB function is linked to completion and
stabilization of adaptive cardiac hypertrophy in response to exercise. Am J Physiol Heart Circ
Physiol 293: H246–H259.
Wei L, Hanna AD, Beard NA, Dulhunty AF. 2009. Unique isoform-specific properties of calsequestrin
in the heart and skeletal muscle. Cell Calcium 45: 474–484.
Wenz T, Rossi SG, Rotundo RL, Spiegelman BM, Moraes CT. 2009. Increased muscle PGC-1α
expression protects from sarcopenia and metabolic disease during aging. Proc Natl Acad Sci 106:
20405–20410.
Wettschureck N, Rutten H, Zywietz A, Gehring D, Wilkie TM, Chen J, Chien KR, Offermanns S.
2001. Absence of pressure overload induced myocardial hypertrophy after conditional inactivation
of Gαq/Gα11 in cardiomyocytes. Nat Med 7: 1236–1240.
Zhang WC, Peng YJ, Zhang GS, He WQ, Qiao YN, Dong YY, Gao YQ, Chen C, Zhang CH, Li W, et
al. 2010. Myosin light chain kinase is necessary for tonic airway smooth muscle contraction. J Biol
Chem 285: 5522–5531.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a006023
CHAPTER 14
D. Grahame Hardie
College of Life Sciences, University of Dundee, Dundee DD1 5EH, Scotland,
United Kingdom
Correspondence: d.g.hardie@dundee.ac.uk
SUMMARY
1 INTRODUCTION
Heterotrophic organisms, including mammals, gain energy from the ingestion
and breakdown (catabolism) of reduced carbon compounds, mainly
carbohydrates, fats, and proteins. A large proportion of the energy released,
rather than appearing simply as heat, is used to convert ADP and inorganic
phosphate (Pi) into ATP. The high intracellular ratio of ATP to ADP thus
created is analogous to the fully charged state of a rechargeable battery,
representing a store of energy that can be used to drive energy-requiring
processes, including the anabolic pathways required for cell maintenance and
growth. Individual cells must constantly adjust their rates of nutrient uptake
and catabolism to balance their rate of ATP consumption, so that they can
maintain a constant high ratio of ATP to ADP. The main control mechanism
used to achieve this energy homeostasis is AMP-activated protein kinase
(AMPK) (see Box 1) (Hardie 2011).
Figure 4. Acute regulation of glycolysis and gluconeogenesis in the liver. Reaction steps unique to
glucose release via gluconeogenesis or glycogenolysis are shown in blue. The steps opposing liver
pyruvate kinase (L-PK) in gluconeogenesis are not shown in detail. During fasting or starvation,
epinephrine and glucagon increase calcium and cyclic AMP (cAMP) levels, activating phosphorylase
kinase and cAMP-dependent protein kinase (PKA), which act together to promote glycogen
breakdown. During starvation, PKA also phosphorylates L-PK and 6-phosphofructo-2-kinase/fructose
2,6-bisphosphatase (PFK2/F2BPase), inactivating the former, and inhibiting the kinase and activating
the phosphatase activity of the latter. This causes a drop in fructose 2,6-bisphosphate levels, which
triggers a net switch from glycolysis to gluconeogenesis.
In the liver, glycolysis is most active in the fed state and is an anabolic
pathway, because it provides precursors for lipid biosynthesis. The liver is
also the major site of gluconeogenesis, the synthesis of glucose from
noncarbohydrate precursors, which is essentially a reversal of glycolysis
except for three irreversible steps in which different reactions are used (blue
arrows in Fig. 4). Gluconeogenesis becomes particularly important as a
source of glucose during starvation, particularly for the brain, which cannot
use fatty acids. The liver therefore must have mechanisms to trigger a switch
from glycolysis to gluconeogenesis during the transition from the fed to the
starved state. A key mediator of this switch is a metabolite that has a purely
regulatory role, fructose 2-6-bisphosphate, which is synthesized and broken
down to fructose 6-phosphate by distinct domains of a single bienzyme
polypeptide termed 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase
(PFK2/FBPase) (Fig. 4).
A key step in glycolysis is the conversion of fructose 6-phosphate to
fructose 1,6-bisphosphate, catalyzed by 6-phosphofructo-1-kinase (PFK1).
PFK1 is allosterically activated by fructose 2, 6-bisphosphate, which also
inhibits the opposing reaction in gluconeogenesis, catalyzed by fructose-1,6-
bisphosphatase. On the transition from the fed to the starved state, glucagon
is released, increasing cyclic AMP levels in the liver and activating PKA.
PKA phosphorylates the liver isoform of PFK2/FBPase, inhibiting its kinase
and activating its phosphatase activity (Rider et al. 2004). The consequent
drop in fructose 2,6-bisphosphate levels both reduces PFK1 activation and
relieves inhibition of fructose-1,6-bisphosphatase, causing a net switch from
glycolysis to gluconeogenesis. In addition, PKA phosphorylates and
inactivates the liver isoform of pyruvate kinase (L-PK) (Riou et al. 1978),
causing additional inhibition of glycolysis at a later step (Fig. 4).
On returning to the fed state again, blood glucose increases, glucagon
levels decrease, and the effects just described are reversed. Some of the
increased flux of glucose into the liver caused by the high blood glucose
levels also enters the pentose phosphate pathway, generating the intermediate
xylulose 5-phosphate. Xylulose 5-phosphate has been found to activate a
protein phosphatase that dephosphorylates PFK2/FBPase, thus switching it
back to the state that favors fructose 2,6-bisphosphate synthesis (Nishimura et
al. 1994).
In muscle, where gluconeogenesis is absent and glycolysis has a purely
catabolic role, it would not make sense for hormones that increase cyclic
AMP (such as epinephrine) to inhibit glycolysis. Indeed, muscle expresses
different isoforms of pyruvate kinase and PFK2/FBPase, which lack the PKA
sites.
8 LIVER—LONG-TERM REGULATION OF
GLUCONEOGENESIS VIA EFFECTS ON GENE
EXPRESSION
Another important tier of regulation of glycolysis and gluconeogenesis occurs
at the level of transcription. Although expression of most genes involved in
these pathways is regulated, research has particularly focused on the genes
encoding the catalytic subunit of glucose-6-phosphatase (G6Pc), and
phosphoenolpyruvate carboxykinase (PEPCK) (Yabaluri and Bashyam
2010). Although often referred to as “gluconeogenic genes,” in fact neither is
involved exclusively with that pathway. Thus, glucose-6-phosphatase
releases into the bloodstream glucose derived from glycogen breakdown as
well as gluconeogenesis (Fig. 4), whereas phosphoenolpyruvate produced by
PEPCK is used as a precursor for biosynthesis of products other than glucose,
including glycerol 3-phosphate used in triglyceride synthesis.
Three important hormonal regulators of transcription of these genes are
glucocorticoids and glucagon (which are released during fasting or starvation
and increase transcription) and insulin (which is released after carbohydrate
feeding and represses transcription). The promoters for these genes contain
hormone response units that bind the critical transcription factors and are
most well defined in the case of the G6Pc promoter (Fig. 5).
Figure 5. Regulation of the G6Pc promoter, showing the approximate location of elements binding the
key transcription factors. Glucocorticoids such as cortisol, in complex with the glucocorticoid receptor
(GR), bind to three sites within the glucocorticoid response unit, enhancing transcription. Cyclic AMP-
dependent protein kinase (PKA) phosphorylates cyclic AMP response element binding protein (CREB),
recruiting CREB-binding protein (CBP), and activating transcription. Finally, Akt phosphorylates
FoxO at multiple sites, triggering the binding of 14-3-3 proteins and their nuclear exclusion, thus
inhibiting transcription. Not shown are the roles of coactivators other than CBP described in the text,
i.e., PGC-1α and CRTC2.
Figure 6. Regulation of processing of SREBPs. The precursor forms of SREBPs bind to the membrane
protein SCAP through interactions between their carboxy-terminal regulatory domain (RD) and the
WD repeat domain (WDD) of SCAP. The SREBP-SCAP complex is retained in the endoplasmic
reticulum by interaction with Insigs. Reduced binding of sterols to the sterol-binding domain (SBD) of
Insig1 and the sterol sensor domain (SSD) of SCAP causes their dissociation, and the SCAP-SREBP2
complex then translocates to the Golgi, where the site 1 and site 2 proteases (S1P and S2P) cleave
SREBP2, releasing the transcription factor domain (TFD) that translocates to the nucleus. Regulation of
SREB1c is similar, except that there appears to be multiple mechanisms that trigger its release from the
ER, including insulin-induced degradation of Insig2.
10 ADIPOCYTES—REGULATION OF FATTY
ACID METABOLISM
Although subcutaneous fat provides thermal insulation, the main metabolic
function of adipocytes is to store fatty acids as triglycerides, neutral lipids
that are very insoluble in water and are deposited in lipid droplets. White
adipocytes, unlike other cells, have a single central triglyceride droplet that
occupies almost the entire volume of the cell. The phospholipid monolayer
that forms its cytoplasmic face is lined with a protein called peripilin1
(Brasaemle 2007). To release fatty acids back into the circulation, the
triglycerides in the lipid droplet must be hydrolyzed back to free fatty acids
(lipolysis). Three lipases are involved, which remove the first, second, and
third fatty acids: (1) adipose tissue triglyceride lipase (ATGL), (2) hormone-
sensitive lipase (HSL), and (3) monacylglycerol lipase. Lipolysis is greatly
enhanced during fasting by glucagon and/or epinephrine, acting via increases
in cyclic AMP; insulin opposes this because Akt phosphorylates and activates
the cyclic AMP phosphodiesterase PDE3B, thus lowering cyclic AMP levels
(Berggreen et al. 2009). HSL is directly phosphorylated and activated by
PKA, although the effect on activity is modest (about twofold) compared
with the effects on lipolysis (at least 100-fold). Phosphorylation of HSL also
triggers its translocation from the cytoplasm to the lipid droplet, thus
increasing its accessibility to substrate (Clifford et al. 2000). However, some
hormone-stimulated release of fatty acids still occurs even in adipocytes from
HSL-deficient mice (Haemmerle et al. 2002). This may be because perilipin1,
which seems to regulate access of lipolytic enzymes to the surface of the lipid
droplet, is also phosphorylated by PKA (Clifford et al. 2000). Adipocytes
from perilipin1 knockout mice have a high basal lipolytic rate that is only
marginally stimulated by cyclic-AMP-elevating agents (Tansey et al. 2001).
The crucial effect of PKA may therefore be to phosphorylate perilipin1,
which alters the accessibility of triglycerides within the lipid droplet to both
ATGL and HSL.
Activation of AMPK opposes the effects of epinephrine and glucagon on
lipolysis, in part because it phosphorylates HSL at sites close to the PKA
sites, antagonizing the activation and translocation induced by PKA (Daval et
al. 2005). Whether AMPK also antagonizes the effect of phosphorylation of
perilipin1 by PKA remains unclear. It might appear paradoxical that AMPK
inhibits lipolysis, because fatty acids are an excellent fuel for oxidative
catabolism. However, the fatty acids produced by lipolysis are not usually
oxidized within the adipocyte, but are released for use elsewhere. If the fatty
acids generated by lipolysis are not rapidly removed from adipocytes either
by export or by oxidative metabolism, they are recycled into triglycerides, an
energy-intensive process in which two molecules of ATP are consumed per
fatty acid. Thus, inhibition of lipolysis by AMPK may ensure that the rate of
lipolysis does not exceed the rate at which the fatty acids can be removed
from the system.
11 BROWN ADIPOCYTES—REGULATION OF
FATTY ACID OXIDATION AND HEAT
PRODUCTION
Brown adipocytes are so called because, unlike white fat cells, they have
abundant mitochondria containing cytochromes that produce their
characteristic color. Increases in cyclic AMP levels induced by epinephrine
trigger lipolysis as in white fat cells, but brown adipocytes differ in that the
fatty acids are not released but are oxidized within their own mitochondria.
Another unique feature of brown adipocytes is that they express uncoupling
protein1 (UCP1), which dissipates the electrochemical gradient produced by
pumping of protons across the inner mitochondrial membrane by the
respiratory chain. The energy expended in brown fat cells therefore mainly
appears in the form of heat, rather than ATP. This heat-generating system is
particularly important in neonatal animals (including humans), but also
occurs in adult rodents exposed to cold environments. Once thought to be
absent in adult humans, improved methodology has shown that brown fat
does indeed occur (van Marken Lichtenbelt et al. 2009). These findings have
rekindled interest in the idea that regulation of energy expenditure by brown
fat might be a way of controlling obesity.
13 CONCLUDING REMARKS
Energy balance in multicellular organisms involves a complex interplay
between energy-utilizing tissues such as muscle, energy-storing tissues such
as adipose tissue, and organs involved in metabolic coordination such as the
liver. These tissues signal to each other via hormones and cytokines that are
either secreted by specialized endocrine cells (e.g., the hypothalamus, islets
of Langerhans, pituitary, thyroid, and adrenal glands) or by the tissues
themselves (e.g., adipokines, released by adipocytes). These hormones either
act at receptors that switch on protein kinase signaling cascades triggered by
second messengers such as PIP3 (insulin), calcium (epinephrine acting at α1
receptors) or cyclic AMP (glucagon), or bind to nuclear receptors that are
transcription factors (cortisol and T3). These signaling cascades interact with
other signaling pathways involved in regulating energy balance at the cell-
autonomous level (e.g., AMPK). The net effect is modulation of carbohydrate
and lipid metabolism, both by direct phosphorylation of metabolic enzymes
and by effects of gene expression or protein turnover. One important
remaining challenge is to understand how cells monitor their levels of energy
reserves such as triglyceride, a process likely to be important in
understanding disorders such as obesity and type 2 diabetes.
REFERENCES
*Reference is in this book.
Baur JA, Pearson KJ, Price NL, Jamieson HA, Lerin C, Kalra A, Prabhu VV, Allard JS, Lopez-Lluch
G, Lewis K, et al. 2006. Resveratrol improves health and survival of mice on a high-calorie diet.
Nature 444: 337–342.
Berggreen C, Gormand A, Omar B, Degerman E, Goransson O. 2009. Protein kinase B activity is
required for the effects of insulin on lipid metabolism in adipocytes. Am J Physiol Endocrinol
Metab 296: E635–E646.
Bonen A, Han XX, Habets DD, Febbraio M, Glatz JF, Luiken JJ. 2007. A null mutation in skeletal
muscle FAT/CD36 reveals its essential role in insulin- and AICAR-stimulated fatty acid
metabolism. Am J Physiol Endocrinol Metab 292: E1740–E1749.
Bouskila M, Hunter RW, Ibrahim AF, Delattre L, Peggie M, van Diepen JA, Voshol PJ, Jensen J,
Sakamoto K. 2010. Allosteric regulation of glycogen synthase controls glycogen synthesis in
muscle. Cell Metab 12: 456–466.
Brasaemle DL. 2007. Thematic review series: Adipocyte biology. The perilipin family of structural
lipid droplet proteins: Stabilization of lipid droplets and control of lipolysis. J Lipid Res 48: 2547–
2559.
Brunet A, Bonni A, Zigmond MJ, Lin MZ, Juo P, Hu LS, Anderson MJ, Arden KC, Blenis J,
Greenberg ME. 1999. Akt promotes cell survival by phosphorylating and inhibiting a Forkhead
transcription factor. Cell 96: 857–868.
Canto C, Jiang LQ, Deshmukh AS, Mataki C, Coste A, Lagouge M, Zierath JR, Auwerx J. 2010.
Interdependence of AMPK and SIRT1 for metabolic adaptation to fasting and exercise in skeletal
muscle. Cell Metab 11: 213–219.
Chen S, Wasserman DH, MacKintosh C, Sakamoto K. 2011. Mice with AS160/TBC1D4-Thr649Ala
knockin mutation are glucose intolerant with reduced insulin sensitivity and altered GLUT4
trafficking. Cell Metab 13: 68–79.
Clarke PR, Hardie DG. 1990. Regulation of HMG-CoA reductase: Identification of the site
phosphorylated by the AMP-activated protein kinase in vitro and in intact rat liver. EMBO J 9:
2439–2446.
Clifford GM, Londos C, Kraemer FB, Vernon RG, Yeaman SJ. 2000. Translocation of hormone-
sensitive lipase and perilipin upon lipolytic stimulation of rat adipocytes. J Biol Chem 275: 5011–
5015.
Daval M, Diot-Dupuy F, Bazin R, Hainault I, Viollet B, Vaulont S, Hajduch E, Ferre P, Foufelle F.
2005. Anti-lipolytic action of AMP-activated protein kinase in rodent adipocytes. J Biol Chem 280:
25250–25257.
Fischer EH, Krebs EG. 1989. Commentary on “The phosphorylase b to a converting enzyme of rabbit
skeletal muscle”. Biochim Biophys Acta 1000: 297–301.
Foretz M, Hebrard S, Leclerc J, Zarrinpashneh E, Soty M, Mithieux G, Sakamoto K, Andreelli F,
Viollet B. 2010. Metformin inhibits hepatic gluconeogenesis in mice independently of the
LKB1/AMPK pathway via a decrease in hepatic energy state. J Clin Invest 120: 2355–2369.
Friedman JM, Halaas JL. 1998. Leptin and the regulation of body weight in mammals. Nature 395:
763–770.
Garg A. 2004. Acquired and inherited lipodystrophies. New Engl J Med 350: 1220–1234.
Greer EL, Dowlatshahi D, Banko MR, Villen J, Hoang K, Blanchard D, Gygi SP, Brunet A. 2007. An
AMPK-FOXO pathway mediates longevity induced by a novel method of dietary restriction in C.
elegans. Curr Biol 17: 1646–1656.
Haemmerle G, Zimmermann R, Hayn M, Theussl C, Waeg G, Wagner E, Sattler W, Magin TM,
Wagner EF, Zechner R. 2002. Hormone-sensitive lipase deficiency in mice causes diglyceride
accumulation in adipose tissue, muscle, and testis. J Biol Chem 277: 4806–4815.
Hardie DG. 2007. AMP-activated/SNF1 protein kinases: Conserved guardians of cellular energy. Nat
Rev Mol Cell Biol 8: 774–785.
Hardie DG. 2011. AMP-activated protein kinase—An energy sensor that regulates all aspects of cell
function. Genes Dev 25: 1895–1908.
Hardie DG, Carling D, Gamblin SJ. 2011. AMP-activated protein kinase: Also regulated by ADP?
Trends Biochem Sci 36: 470–477.
* Harrison DA. 2012. The JAK/STAT pathway. Cold Spring Harb Perspect Biol 4: a011205.
Hawley SA, Ross FA, Chevtzoff C, Green KA, Evans A, Fogarty S, Towler MC, Brown LJ, Ogunbayo
OA, Evans AM, et al. 2010. Use of cells expressing γ subunit variants to identify diverse
mechanisms of AMPK activation. Cell Metab 11: 554–565.
* Heldin CH, Lu B, Evans R, Gutkind JS. 2014. Signals and receptors. Cold Spring Harb Perspect Biol
doi: 10.1101/cshperspect.a005900.
* Hemmings BA, Restuccia DF. 2012. The PI3K-PKB/Akt pathway. Cold Spring Harb Perspect Biol
4: a011189.
Hu FB, Manson JE, Stampfer MJ, Colditz G, Liu S, Solomon CG, Willett WC. 2001. Diet, lifestyle,
and the risk of type 2 diabetes mellitus in women. New Engl J Med 345: 790–797.
Jager S, Handschin C, St-Pierre J, Spiegelman BM. 2007. AMP-activated protein kinase (AMPK)
action in skeletal muscle via direct phosphorylation of PGC-1α. Proc Natl Acad Sci 104: 12017–
12022.
Jorgensen SB, Nielsen JN, Birk JB, Olsen GS, Viollet B, Andreelli F, Schjerling P, Vaulont S, Hardie
DG, Hansen BF, et al. 2004. The α2-5’AMP-activated protein kinase is a site 2 glycogen synthase
kinase in skeletal muscle and is responsive to glucose loading. Diabetes 53: 3074–3081.
Kabashima T, Kawaguchi T, Wadzinski BE, Uyeda K. 2003. Xylulose 5-phosphate mediates glucose-
induced lipogenesis by xylulose 5-phosphate-activated protein phosphatase in rat liver. Proc Natl
Acad Sci 100: 5107–5112.
Kadowaki T, Yamauchi T. 2005. Adiponectin and adiponectin receptors. Endocr Rev 26: 439–451.
Kato S, Ding J, Pisck E, Jhala US, Du K. 2008. COP1 functions as a FoxO1 ubiquitin E3 ligase to
regulate FoxO1-mediated gene expression. J Biol Chem 283: 35464–35473.
Koh HJ, Toyoda T, Fujii N, Jung MM, Rathod A, Middelbeek RJ, Lessard SJ, Treebak JT, Tsuchihara
K, Esumi H, et al. 2010. Sucrose nonfermenting AMPK-related kinase (SNARK) mediates
contraction-stimulated glucose transport in mouse skeletal muscle. Proc Natl Acad Sci 107: 15541–
15546.
Kubota N, Yano W, Kubota T, Yamauchi T, Itoh S, Kumagai H, Kozono H, Takamoto I, Okamoto S,
Shiuchi T, et al. 2007. Adiponectin stimulates AMP-activated protein kinase in the hypothalamus
and increases food intake. Cell Metab 6: 55–68.
* Kuo IY, Ehrlich BE. 2014. Signaling in muscle contraction. Cold Spring Harb Perspect Biol doi:
10.1101/cshperspect.a006023.
Lin J, Handschin C, Spiegelman BM. 2005. Metabolic control through the PGC-1 family of
transcription coactivators. Cell Metab 1: 361–370.
Lopez M, Varela L, Vazquez MJ, Rodriguez-Cuenca S, Gonzalez CR, Velagapudi VR, Morgan DA,
Schoenmakers E, Agassandian K, Lage R, et al. 2010. Hypothalamic AMPK and fatty acid
metabolism mediate thyroid regulation of energy balance. Nat Med 16: 1001–1008.
Mair W, Morantte I, Rodrigues AP, Manning G, Montminy M, Shaw RJ, Dillin A. 2011. Lifespan
extension induced by AMPK and calcineurin is mediated by CRTC-1 and CREB. Nature 470: 404–
408.
McCrimmon RJ, Shaw M, Fan X, Cheng H, Ding Y, Vella MC, Zhou L, McNay EC, Sherwin RS.
2008. Key role for AMP-activated protein kinase in the ventromedial hypothalamus in regulating
counterregulatory hormone responses to acute hypoglycemia. Diabetes 57: 444–450.
McManus EJ, Sakamoto K, Armit LJ, Ronaldson L, Shpiro N, Marquez R, Alessi DR. 2005. Role that
phosphorylation of GSK3 plays in insulin and Wnt signalling defined by knockin analysis. EMBO J
24: 1571–1583.
Merrill GM, Kurth E, Hardie DG, Winder WW. 1997. AICAR decreases malonyl-CoA and increases
fatty acid oxidation in skeletal muscle of the rat. Am J Physiol 273: E1107–E1112.
Munday MR, Campbell DG, Carling D, Hardie DG. 1988. Identification by amino acid sequencing of
three major regulatory phosphorylation sites on rat acetyl-CoA carboxylase. Eur J Biochem 175:
331–338.
Nishimura M, Fedorov S, Uyeda K. 1994. Glucose-stimulated synthesis of fructose 2,6-bisphosphate in
rat liver. Dephosphorylation of fructose 6-phosphate, 2-kinase:fructose 2,6-bisphosphatase and
activation by a sugar phosphate. J Biol Chem 269: 26100–26106.
O’Neill HM, Maarbjerg SJ, Crane JD, Jeppesen J, Jorgensen SB, Schertzer JD, Shyroka O, Kiens B,
van Denderen BJ, Tarnopolsky MA, et al. 2011. AMP-activated protein kinase (AMPK) β1β2
muscle null mice reveal an essential role for AMPK in maintaining mitochondrial content and
glucose uptake during exercise. Proc Natl Acad Sci 108: 16092–16097.
Oakhill JS, Steel R, Chen ZP, Scott JW, Ling N, Tam S, Kemp BE. 2011. AMPK is a direct adenylate
charge-regulated protein kinase. Science 332: 1433–1435.
Ouyang J, Parakhia RA, Ochs RS. 2011. Metformin activates AMP kinase through inhibition of AMP
deaminase. J Biol Chem 286: 1–11.
Owen MR, Doran E, Halestrap AP. 2000. Evidence that metformin exerts its anti-diabetic effects
through inhibition of complex 1 of the mitochondrial respiratory chain. Biochem J 348: 607–614.
Raghow R, Yellaturu C, Deng X, Park EA, Elam MB. 2008. SREBPs: The crossroads of physiological
and pathological lipid homeostasis. Trends Endocrinol Metab 19: 65–73.
Ramaiah A, Hathaway JA, Atkinson DE. 1964. Adenylate as a metabolic regulator. Effect on yeast
phosphofructokinase kinetics. J Biol Chem 239: 3619–3622.
Rider MH, Bertrand L, Vertommen D, Michels PA, Rousseau GG, Hue L. 2004. 6-phosphofructo-2-
kinase/fructose-2,6-bisphosphatase: Head-to-head with a bifunctional enzyme that controls
glycolysis. Biochem J 381: 561–579.
Riou JP, Claus TH, Pilkis SJ. 1978. Stimulation of glucagon of in vivo phosphorylation of rat hepatic
pyruvate kinase. J Biol Chem 253: 656–659.
Sakamoto K, Holman GD. 2008. Emerging role for AS160/TBC1D4 and TBC1D1 in the regulation of
GLUT4 traffic. Am J Physiol Endocrinol Metab 295: E29–E37.
Sakamoto K, McCarthy A, Smith D, Green KA, Hardie DG, Ashworth A, Alessi DR. 2005. Deficiency
of LKB1 in skeletal muscle prevents AMPK activation and glucose uptake during contraction.
EMBO J 24: 1810–1820.
Samuel VT, Petersen KF, Shulman GI. 2010. Lipid-induced insulin resistance: Unravelling the
mechanism. Lancet 375: 2267–2277.
Shetty S, Kusminski CM, Scherer PE. 2009. Adiponectin in health and disease: Evaluation of
adiponectin-targeted drug development strategies. Trends Pharmacol Sci 30: 234–239.
Tansey JT, Sztalryd C, Gruia-Gray J, Roush DL, Zee JV, Gavrilova O, Reitman ML, Deng CX, Li C,
Kimmel AR, et al. 2001. Perilipin ablation results in a lean mouse with aberrant adipocyte lipolysis,
enhanced leptin production, and resistance to diet-induced obesity. Proc Natl Acad Sci 98: 6494–
6499.
Tissenbaum HA, Guarente L. 2001. Increased dosage of a sir-2 gene extends lifespan in
Caenorhabditis elegans. Nature 410: 227–230.
Um JH, Park SJ, Kang H, Yang S, Foretz M, McBurney MW, Kim MK, Viollet B, Chung JH. 2010.
AMP-activated protein kinase-deficient mice are resistant to the metabolic effects of resveratrol.
Diabetes 59: 554–563.
Uyeda K, Repa JJ. 2006. Carbohydrate response element binding protein, ChREBP, a transcription
factor coupling hepatic glucose utilization and lipid synthesis. Cell Metab 4: 107–110.
van Marken Lichtenbelt WD, Vanhommerig JW, Smulders NM, Drossaerts JM, Kemerink GJ, Bouvy
ND, Schrauwen P, Teule GJ. 2009. Cold-activated brown adipose tissue in healthy men. New Engl
J Med 360: 1500–1508.
van Schaftingen E, Gerin I. 2002. The glucose-6-phosphatase system. Biochem J 362: 513–532.
Wende AR, Huss JM, Schaeffer PJ, Giguere V, Kelly DP. 2005. PGC-1α coactivates PDK4 gene
expression via the orphan nuclear receptor ERRα: A mechanism for transcriptional control of
muscle glucose metabolism. Mol Cell Biol 25: 10684–10694.
Xiao B, Sanders MJ, Underwood E, Heath R, Mayer FV, Carmena D, Jing C, Walker PA, Eccleston JF,
Haire LF, et al. 2011. Structure of mammalian AMPK and its regulation by ADP. Nature 472: 230–
233.
Yabaluri N, Bashyam MD. 2010. Hormonal regulation of gluconeogenic gene transcription in the liver.
J Biosci 35: 473–484.
Yang T, Espenshade PJ, Wright ME, Yabe D, Gong Y, Aebersold R, Goldstein JL, Brown MS. 2002.
Crucial step in cholesterol homeostasis: Sterols promote binding of SCAP to INSIG-1, a membrane
protein that facilitates retention of SREBPs in ER. Cell 110: 489–500.
Yang W, Lu J, Weng J, Jia W, Ji L, Xiao J, Shan Z, Liu J, Tian H, Ji Q, et al. 2010. Prevalence of
diabetes among men and women in China. New Engl J Med 362: 1090–1101.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a006031
CHAPTER 15
SUMMARY
Outline
1 Introduction
2 DAMPs and PAMPs trigger the innate immune response
3 Toll-like receptors (TLRs)
4 RIG-I-like receptors (RLRs)
5 Nod-like receptors (NLRs)
6 The proinflammatory cytokine tumor necrosis factor (TNF)
7 Selectins and integrins
8 G-protein-coupled receptors (GPCRs)
9 Fc receptors
10 Inflammation as a risk factor for cancer
11 Concluding remarks
References
1 INTRODUCTION
The role of the inflammatory response is to combat infection and tissue
injury. Innate immune cells residing in tissues, such as macrophages,
fibroblasts, mast cells, and dendritic cells, as well as circulating leukocytes,
including monocytes and neutrophils, recognize pathogen invasion or cell
damage with intracellular or surface-expressed pattern recognition receptors
(PRRs). These receptors detect, either directly or indirectly, pathogen-
associated molecular patterns (PAMPs), such as microbial nucleic acids,
lipoproteins, and carbohydrates, or damage-associated molecular patterns
(DAMPs) released from injured cells. Activated PRRs then oligomerize and
assemble large multisubunit complexes that initiate signaling cascades that
trigger the release of factors that promote recruitment of leukocytes to the
region.
Vascular alterations play an important role in the inflammatory response
(Fig. 1). Histamine, prostaglandins, and nitric oxide act on vascular smooth
muscle to cause vasodilation, which increases blood flow and brings in
circulating leukocytes, whereas inflammatory mediators including histamine
and leukotrienes act on endothelial cells to increase vascular permeability and
allow plasma proteins and leukocytes to exit the circulation. Cytokines such
as tumor necrosis factor (TNF) and interleukin 1 (IL1) promote leukocyte
extravasation by increasing the levels of leukocyte adhesion molecules on
endothelial cells. Activated innate immune cells at the site of infection or
injury, including dendritic cells, macrophages, and neutrophils, remove
foreign particles and host debris by phagocytosis, plus they also secrete
cytokines that shape the slower, lymphocyte-mediated adaptive immune
response.
Figure 1. Cells and mediators of the inflammatory response. Molecules derived from plasma proteins
and cells in response to tissue damage or pathogens mediate inflammation by stimulating vascular
changes, plus leukocyte migration and activation. Granulocytes include neutrophils, basophils, and
eosinophils.
12b Unknown
13b Unknown
The MAPKs JNK and p38α, like TAK1, are activated downstream from
TLR2 and TLR4 in a cIAP-dependent manner. The TAK1-containing
signaling complex translocates into the cytosol and recruits the kinase MKK4
to phosphorylate and activate JNK (Tseng et al. 2010). It probably also
recruits MKK3 and MKK6 to activate p38α because both kinases associate
with TRAF6 in response to LPS (Wan et al. 2009). p38α is required for
activation of the transcription factors CREB and c/EBPβ, and it contributes to
the induction of several genes, including those encoding chemokines (Cxcl1,
Cxcl2), cytokines (IL10, IL12b, IL1a, and IL1b), and regulators of
extracellular matrix remodeling (Mmp13) and cell adhesion (Vcam1) (Kang
et al. 2008; Kim et al. 2008). JNK regulates the activity of the AP1
transcription factor and stimulates expression of proinflammatory mediators
such as TNF (Das et al. 2009).
Activation of the IKK complex is required for NF-κB-dependent
transcription as well as transcriptional responses downstream from the
MAPK ERK. IKKβ substrates include p105, the precursor of the p50 NF-κB1
transcription factor, as well as the IκB proteins that sequester NF-κB
transcription factors in the cytosol. Phosphorylation by IKKβ targets these
substrates for K48-linked polyubiquitylation by the E3 ubiquitin ligase SCFβ-
TrCP and subsequent proteasomal degradation (Kanarek et al. 2010).
Degradation of p105, which exists in a complex with the kinase Tpl2,
activates a Tpl2–MEK1–ERK kinase cascade that leads to the induction of
genes such as Ptgs2 by the CREB/ATF family of transcription factors
(Banerjee and Gerondakis 2007). The cyclooxygenase 2 (COX2) enzyme
encoded by Ptgs2 is involved in the synthesis of prostaglandins, which are
important mediators of pain, inflammation, and fever. IκB degradation allows
dimeric NF-κB transcription factors composed largely of RelA (p65) and NF-
κB1 (p50) subunits to accumulate in the nucleus and drive expression of a
large number of proinflammatory genes (Table 3 lists a subset of these
genes).
Figure 3. Signaling by RIG-I. (A) RIG-I binding to dsRNA that has a 5′ triphosphate and polyubiquitin,
the latter generated by the ubiquitin ligase TRIM25 and E2 ubiquitin-conjugating enzymes Ubc5 and
Ubc13, promotes RIG-I binding to mitochondrial MAVS. Subsequently, a larger complex containing
the adaptor proteins CARD9 and BCL10 is assembled for MAPK and NF-κB activation. TRAF3, the
kinases TBK1 and IKKε, and ER-resident protein STING are required for activation of transcription
factors IRF3 and IRF7. (B) Functional outputs of some of the genes up-regulated by MAVS signaling.
IKKβ activation by TNF triggers not only NF-κB transcription but also
the Tpl2–MEK1–ERK kinase cascade activated by TLRs (see above). In
fibroblasts, but not macrophages or B cells, Tpl2 activation by TNF has been
linked to activation of the MKK4–JNK pathway as well. In addition,
activation of the kinase MSK1 by ERK may enhance NF-κB transcriptional
activity through phosphorylation of RelA (Banerjee and Gerondakis 2007).
Genetic studies indicate that TNF-induced JNK activation is mediated largely
by upstream kinases TAK1 and MKK7, whereas p38 activation requires
TAK1 and MKK3/MKK6 (Brancho et al. 2003; Sato et al. 2005; Shim et al.
2005).
9 Fc RECEPTORS
Repeated exposure to a polyvalent foreign substance can elicit an
inflammatory response called a hypersensitivity reaction if the host makes
antibodies against the substance. Immune complexes containing the antigen
and IgG or IgM antibodies activate complement proteases, culminating in the
generation of C3a and C5a, which signal leukocyte recruitment and activation
(see above); the opsonin C3b, which coats and promotes phagocytosis of
bacteria; and the membrane attack complex for bacterial cell lysis (C5b-9). In
addition, complexes containing IgG or IgE antibodies engage Fc receptors on
leukocytes. Members of the Fc receptor family are type I transmembrane
proteins (with the exception of human GPI-anchored FcγRIIIB) that produce
activating (human FcγRI, FcγRIIA, FcγRIIC, FcγRIIIA, FcγRIIIB, and
FcεRI) or inhibitory (human FcγRIIB) signals. Mast cells expressing the
high-affinity receptor for IgE, FcεRI, play a central role in allergic reactions.
FcεRI engagement causes intracellular granules to fuse with the plasma
membrane such that preformed inflammatory mediators including histamine,
serotonin, and proteases are released into the extracellular environment.
Activated mast cells also secrete proinflammatory prostaglandins,
leukotrienes, and cytokines, but these are synthesized de novo.
9.1 FcεRI
FcεRI is an αβγ2 heterotetramer. Its α-chain contains extracellular Ig-like
domains for binding the heavy-chain constant region of IgE, whereas the β-
chain and a γ-chain homodimer transduce signals via cytoplasmic ITAMs
(Fig. 7) (Ch. 16 [Cantrell 2014]). IgE-induced clustering of FcεRI promotes
activation of Src family kinases Lyn and Fyn. Lyn substrates include both
positive and negative regulators of mast cell activation, which fine-tune the
magnitude and duration of the response. Lyn stimulates activation by
phosphorylating the FcRγ ITAM, which recruits the SH2 domains in Syk.
Subsequent Syk-dependent phosphorylation of the transmembrane adaptors
LAT1 and LAT2 recruits additional SH2-containing signaling components,
such as PLCγ and the adaptors Grb2 and Gads. SH3 domains in Grb2 and
Gads bind proline-rich regions in additional proteins such as the adaptors
SLP76 and Gab2. SLP76 interacts with the Rho/Rac GEF VAV1, which
contributes to PLCγ and JNK activation. Fyn-dependent phosphorylation of
Gab2 recruits the SH2-containing p85 regulatory subunit of PI3Kδ. PIP3
produced by PI3Kδ retains proteins containing plextrin homology (PH)
domains at the plasma membrane, such as PLCγ, Gab2, Akt, and Btk. The
kinase Btk phosphorylates and enhances the activity of PLCγ (Alvarez-Errico
et al. 2009).
Figure 7. Signaling by FcεRI. (A) Binding of the Fc region of antigen-bound IgE to FcεRI activates the
Src family kinases Lyn and Fyn. Tyrosine phosphorylation of the FcRγ ITAM recruits the tyrosine
kinase Syk, which is required for phosphorylation of LAT transmembrane adaptor proteins.
Phosphorylated LAT1 binds PLCγ and the adaptors Gads and Grb2. Gads recruits the adaptor SLP76,
which regulates activation of PLCγ and the GEF Vav1. Grb2 binds Gab2, which is phosphorylated by
Fyn and binds the p85 regulatory subunit of PI3Kδ. PIP3 generated by PI3Kδ retains signaling
components such as Gab2, PLCγ, and Btk at the plasma membrane. IP3 generated by PLCγ depletes
ER calcium stores, which causes a STIM1-dependent influx of calcium that promotes mast cell
degranulation. Elevated intracellular calcium also activates the phosphatase calcineurin, stimulates
NFAT-dependent gene expression, and triggers the translocation of cPLA2 and 5-lipoxygenase (5-LO)
to the nuclear envelope, cytoplasmic lipid bodies, or ER. cPLA2 releases arachidonic acid from
membrane phospholipids. COX enzymes and downstream synthases metabolize arachidonic acid into
prostaglandins and thromboxane, whereas leukotriene (LT) synthesis from arachidonic acid involves
five-lipoxygenase-activating protein (FLAP), 5-LO, and downstream LTC4 synthase or LTA4
hydrolase. DAG generated by PLCγ activates PKC, which is important for IKK activation via MALT1,
BCl10, and TRAF6, as well as subsequent NF-κB-dependent gene transcription. IKKβ has also been
implicated in mast cell degranulation independent of NF-κB activation. (B) Functional outputs of some
of the genes up-regulated by FcεRI signaling.
11 CONCLUDING REMARKS
Many of the major players in inflammatory signaling have been identified,
but the importance and complexity of posttranslational modifications such as
ubiquitylation in these pathways continue to be unraveled. Binding of TLRs,
TNF and IL1 receptors, GPCRs, integrins, selectins, and Fc receptors to their
ligands triggers the formation of multisubunit signaling complexes, but it
remains to be seen how diverse inflammatory stimuli can activate
intracellular PRRs such as NLRP3 and NLRC4. An attractive hypothesis is
that posttranslational modifications to NLR family members are key to their
activation. Once activated, both surface and intracellular PRRs stimulate
transcription of inflammatory genes; TLRs, RLRs, and some NLRs (e.g.,
NOD1 and NOD2) engage common downstream signaling pathways to
stimulate transcription factors such as NF-κB, AP1, CREB, and c/EBPβ,
whereas caspase-1-activating PRRs (e.g., AIM2, NLRP3, and NLRC4)
stimulate similar pathways indirectly via the secretion of IL1β and IL18.
Going forward, it will be important to understand how innate immune cells
exposed to multiple inflammatory mediators and stimuli in vivo integrate
signals from diverse receptors, because this will offer insight into what
critical components might be targeted for therapeutic benefit in inflammatory
disorders.
REFERENCES
*Reference is in this book.
Albrecht EA, Chinnaiyan AM, Varambally S, Kumar-Sinha C, Barrette TR, Sarma JV, Ward PA. 2004.
C5a-induced gene expression in human umbilical vein endothelial cells. Am J Pathol 164: 849–859.
Alvarez SE, Harikumar KB, Hait NC, Allegood J, Strub GM, Kim EY, Maceyka M, Jiang H, Luo C,
Kordula T, et al. 2010. Sphingosine-1-phosphate is a missing cofactor for the E3 ubiquitin ligase
TRAF2. Nature 465: 1084–1088.
Alvarez-Errico D, Lessmann E, Rivera J. 2009. Adapters in the organization of mast cell signaling.
Immunol Rev 232: 195–217.
Andreu P, Johansson M, Affara NI, Pucci F, Tan T, Junankar S, Korets L, Lam J, Tawfik D, DeNardo
DG, et al. 2010. FcRγ activation regulates inflammation-associated squamous carcinogenesis.
Cancer Cell 17: 121–134.
Baba Y, Nishida K, Fujii Y, Hirano T, Hikida M, Kurosaki T. 2008. Essential function for the calcium
sensor STIM1 in mast cell activation and anaphylactic responses. Nat Immunol 9: 81–88.
Balachandran S, Thomas E, Barber GN. 2004. A FADD-dependent innate immune mechanism in
mammalian cells. Nature 432: 401–405.
Banerjee A, Gerondakis S. 2007. Coordinating TLR-activated signaling pathways in cells of the
immune system. Immunol Cell Biol 85: 420–424.
Bertrand MJ, Doiron K, Labbe K, Korneluk RG, Barker PA, Saleh M. 2009. Cellular inhibitors of
apoptosis cIAP1 and cIAP2 are required for innate immunity signaling by the pattern recognition
receptors NOD1 and NOD2. Immunity 30: 789–801.
* Bootman MD. 2012. Calcium signaling. Cold Spring Harb Perspect Biol 4: a011171.
Brancho D, Tanaka N, Jaeschke A, Ventura JJ, Kelkar N, Tanaka Y, Kyuuma M, Takeshita T, Flavell
RA, Davis RJ. 2003. Mechanism of p38 MAP kinase activation in vivo. Genes Dev 17: 1969–1978.
Brechard S, Plancon S, Melchior C, Tschirhart EJ. 2009. STIM1 but not STIM2 is an essential
regulator of Ca2+ influx-mediated NADPH oxidase activity in neutrophil-like HL-60 cells.
Biochem Pharmacol 78: 504–513.
Broz P, Newton K, Lamkanfi M, Mariathasan S, Dixit VM, Monack DM. 2010. Redundant roles for
inflammasome receptors NLRP3 and NLRC4 in host defense against Salmonella. J Exp Med 207:
1745–1755.
Camps M, Carozzi A, Schnabel P, Scheer A, Parker PJ, Gierschik P. 1992. Isozyme-selective
stimulation of phospholipase C-β2 by G protein β γ-subunits. Nature 360: 684–686.
* Cantrell D. 2014. Signaling in lymphocyte activation. Cold Spring Harb Perspect Biol doi:
10.1101/cshperspect.a018788.
Chang M, Jin W, Sun SC. 2009. Peli1 facilitates TRIF-dependent Toll-like receptor signaling and
proinflammatory cytokine production. Nat Immunol 10: 1089–1095.
Chen Y, Pappu BP, Zeng H, Xue L, Morris SW, Lin X, Wen R, Wang D. 2007. B cell lymphoma 10 is
essential for FcεR-mediated degranulation and IL-6 production in mast cells. J Immunol 178: 49–
57.
Chen NJ, Chio II, Lin WJ, Duncan G, Chau H, Katz D, Huang HL, Pike KA, Hao Z, Su YW, et al.
2008. Beyond tumor necrosis factor receptor: TRADD signaling in toll-like receptors. Proc Natl
Acad Sci 105: 12429–12434.
Cho YS, Challa S, Moquin D, Genga R, Ray TD, Guildford M, Chan FK. 2009. Phosphorylation-
driven assembly of the RIP1–RIP3 complex regulates programmed necrosis and virus-induced
inflammation. Cell 137: 1112–1123.
Colvin RA, Means TK, Diefenbach TJ, Moita LF, Friday RP, Sever S, Campanella GS, Abrazinski T,
Manice LA, Moita C, et al. 2010. Synaptotagmin-mediated vesicle fusion regulates cell migration.
Nat Immunol 11: 495–502.
Condliffe AM, Webb LM, Ferguson GJ, Davidson K, Turner M, Vigorito E, Manifava M, Chilvers ER,
Stephens LR, Hawkins PT. 2006. RhoG regulates the neutrophil NADPH oxidase. J Immunol 176:
5314–5320.
Cusson-Hermance N, Khurana S, Lee TH, Fitzgerald KA, Kelliher MA. 2005. Rip1 mediates the Trif-
dependent toll-like receptor 3- and 4-induced NF-κB activation but does not contribute to interferon
regulatory factor 3 activation. J Biol Chem 280: 36560–36566.
Das M, Sabio G, Jiang F, Rincon M, Flavell RA, Davis RJ. 2009. Induction of hepatitis by JNK-
mediated expression of TNF-α. Cell 136: 249–260.
* Devreotes P, Horwitz AR. 2014. Signaling networks that regulate cell migration. Cold Spring Harb
Perspect Biol doi: 10.1101/cshperspect.a005959.
Döffinger R, Smahi A, Bessia C, Geissmann F, Feinberg J, Durandy A, Bodemer C, Kenwrick S,
Dupuis-Girod S, Blanche S, et al. 2001. X-linked anhidrotic ectodermal dysplasia with
immunodeficiency is caused by impaired NF-κB signaling. Nat Genet 27: 277–285.
Ermolaeva MA, Michallet MC, Papadopoulou N, Utermohlen O, Kranidioti K, Kollias G, Tschopp J,
Pasparakis M. 2008. Function of TRADD in tumor necrosis factor receptor 1 signaling and in
TRIF-dependent inflammatory responses. Nat Immunol 9: 1037–1046.
Ferguson GJ, Milne L, Kulkarni S, Sasaki T, Walker S, Andrews S, Crabbe T, Finan P, Jones G,
Jackson S, et al. 2007. PI(3)Kγ has an important context-dependent role in neutrophil chemokinesis.
Nat Cell Biol 9: 86–91.
Fujishima H, Sanchez Mejia RO, Bingham CO III, Lam BK, Sapirstein A, Bonventre JV, Austen KF,
Arm JP. 1999. Cytosolic phospholipase A2 is essential for both the immediate and the delayed
phases of eicosanoid generation in mouse bone marrow-derived mast cells. Proc Natl Acad Sci 96:
4803–4807.
Gerlach B, Cordier SM, Schmukle AC, Emmerich CH, Rieser E, Haas TL, Webb AI, Rickard JA,
Anderton H, Wong WW, et al. 2011. Linear ubiquitination prevents inflammation and regulates
immune signalling. Nature 471: 591–596.
* Green DR, Llambi F. 2014. Cell death signaling. Cold Spring Harb Perspect Biol doi:
10.1101/cshperspect.a006080.
Grivennikov S, Karin E, Terzic J, Mucida D, Yu GY, Vallabhapurapu S, Scheller J, Rose-John S,
Cheroutre H, Eckmann L, et al. 2009. IL-6 and Stat3 are required for survival of intestinal epithelial
cells and development of colitis-associated cancer. Cancer Cell 15: 103–113.
Gu H, Botelho RJ, Yu M, Grinstein S, Neel BG. 2003. Critical role for scaffolding adapter Gab2 in
FcγR-mediated phagocytosis. J Cell Biol 161: 1151–1161.
Guerra C, Collado M, Navas C, Schuhmacher AJ, Hernández-Porras I, Cañamero M, Rodriguez-Justo
M, Serrano M, Barbacid M. 2011. Pancreatitis-induced inflammation contributes to pancreatic
cancer by inhibiting oncogene-induced senescence. Cancer Cell 19: 728–739.
Haas TL, Emmerich CH, Gerlach B, Schmukle AC, Cordier SM, Rieser E, Feltham R, Vince J,
Warnken U, Wenger T, et al. 2009. Recruitment of the linear ubiquitin chain assembly complex
stabilizes the TNF-R1 signaling complex and is required for TNF-mediated gene induction. Mol
Cell 36: 831–844.
Hall AB, Gakidis MA, Glogauer M, Wilsbacher JL, Gao S, Swat W, Brugge JS. 2006. Requirements
for Vav guanine nucleotide exchange factors and Rho GTPases in FcγR- and complement-mediated
phagocytosis. Immunity 24: 305–316.
He S, Wang L, Miao L, Wang T, Du F, Zhao L, Wang X. 2009. Receptor interacting protein kinase-3
determines cellular necrotic response to TNF-α. Cell 137: 1100–1111.
Hitotsumatsu O, Ahmad RC, Tavares R, Wang M, Philpott D, Turer EE, Lee BL, Advincula R, Malynn
BA, Werts C, et al. 2008. The ubiquitin-editing enzyme A20 restricts nucleotide-binding
oligomerization domain containing 2-triggered signals. Immunity 28: 381–390.
Hornung V, Latz E. 2010. Intracellular DNA recognition. Nat Rev Immunol 10: 123–130.
Hoshino K, Sasaki I, Sugiyama T, Yano T, Yamazaki C, Yasui T, Kikutani H, Kaisho T. 2010. Critical
role of IκB kinase α in TLR7/9-induced type I IFN production by conventional dendritic cells. J
Immunol 184: 3341–3345.
Hou F, Sun L, Zheng H, Skaug B, Jiang Q, Chen ZJ. 2011. MAVS forms functional prion-like
aggregates to activate and propagate antiviral innate immune response. Cell 146: 448–461.
Ikeda F, Deribe YL, Skånland SS, Stieglitz B, Grabbe C, Franz-Wachtel M, van Wijk SJ, Goswami P,
Nagy V, Terzic J, et al. 2011. SHARPIN forms a linear ubiquitin ligase complex regulating NF-κB
activity and apoptosis. Nature 471: 637–641.
Ishikawa H, Barber GN. 2008. STING is an endoplasmic reticulum adaptor that facilitates innate
immune signalling. Nature 455: 674–678.
Israel A. 2010. The IKK complex, a central regulator of NF-κB activation. Cold Spring Harb Perspect
Biol 2: a000158.
Jakus Z, Simon E, Frommhold D, Sperandio M, Mocsai A. 2009. Critical role of phospholipase Cγ2 in
integrin and Fc receptor-mediated neutrophil functions and the effector phase of autoimmune
arthritis. J Exp Med 206: 577–593.
Jiang F, Ramanathan A, Miller MT, Tang CQ, Gale M, Patel SS, Marcotrigiano J. 2011. Structural
basis of RNA recognition and activation by innate immune receptor RIG-I. Nature 479: 423–427.
Kanarek N, London N, Schueler-Furman O, Ben-Neriah Y. 2010. Ubiquitination and degradation of the
inhibitors of NF-κB. Cold Spring Harb Perspect Biol 2: a000166.
Kang YJ, Chen J, Otsuka M, Mols J, Ren S, Wang Y, Han J. 2008. Macrophage deletion of p38α
partially impairs lipopolysaccharide-induced cellular activation. J Immunol 180: 5075–5082.
Kayagaki N, Wong MT, Stowe IB, Ramani SR, Gonzalez LC, Akashi-Takamura S, Miyake K, Zhang
J, Lee WP, Muszynski A, et al. 2013. Noncanonical inflammasome activation by intracellular LPS
independent of TLR4. Science doi: 10.1126/science.1240248.
Kim C, Sano Y, Todorova K, Carlson BA, Arpa L, Celada A, Lawrence T, Otsu K, Brissette JL, Arthur
JS, et al. 2008. The kinase p38α serves cell type-specific inflammatory functions in skin injury and
coordinates pro- and anti-inflammatory gene expression. Nat Immunol 9: 1019–1027.
Klemm S, Gutermuth J, Hultner L, Sparwasser T, Behrendt H, Peschel C, Mak TW, Jakob T, Ruland J.
2006. The Bcl10–Malt1 complex segregates FcεRI-mediated nuclear factor κB activation and
cytokine production from mast cell degranulation. J Exp Med 203: 337–347.
Komander D, Reyes-Turcu F, Licchesi JD, Odenwaelder P, Wilkinson KD, Barford D. 2009. Molecular
discrimination of structurally equivalent Lys 63-linked and linear polyubiquitin chains. EMBO Rep
10: 466–473.
Kunisaki Y, Nishikimi A, Tanaka Y, Takii R, Noda M, Inayoshi A, Watanabe K, Sanematsu F,
Sasazuki T, Sasaki T, et al. 2006. DOCK2 is a Rac activator that regulates motility and polarity
during neutrophil chemotaxis. J Cell Biol 174: 647–652.
Laplantine E, Fontan E, Chiaravalli J, Lopez T, Lakisic G, Veron M, Agou F, Israel A. 2009. NEMO
specifically recognizes K63-linked poly-ubiquitin chains through a new bipartite ubiquitin-binding
domain. EMBO J 28: 2885–2895.
Li Z, Jiang H, Xie W, Zhang Z, Smrcka AV, Wu D. 2000. Roles of PLC-β2 and -β3 and PI3Kγ in
chemoattractant-mediated signal transduction. Science 287: 1046–1049.
Li Z, Hannigan M, Mo Z, Liu B, Lu W, Wu Y, Smrcka AV, Wu G, Li L, Liu M, Huang CK, Wu D.
2003. Directional sensing requires Gβ γ-mediated PAK1 and PIX α-dependent activation of Cdc42.
Cell 114: 215–227.
Li S, Wang L, Dorf ME. 2009. PKC phosphorylation of TRAF2 mediates IKKα/β recruitment and
K63-linked polyubiquitination. Mol Cell 33: 30–42.
Lin SC, Lo YC, Wu H. 2010. Helical assembly in the MyD88–IRAK4–IRAK2 complex in TLR/IL-1R
signalling. Nature 465: 885–890.
Lowell CA. 2011. Src-family and Syk kinases in activating and inhibitory pathways in innate immune
cells: Signaling cross talk. Cold Spring Harb Perspect Biol 3: a002352.
Lu C, Xu H, Ranjith-Kumar CT, Brooks MT, Hou TY, Hu F, Herr AB, Strong RK, Kao CC, Li P.
2010. The structural basis of 5′ triphosphate double-stranded RNA recognition by RIG-I C-terminal
domain. Structure 18: 1032–1043.
Mariathasan S, Newton K, Monack DM, Vucic D, French DM, Lee WP, Roose-Girma M, Erickson S,
Dixit VM. 2004. Differential activation of the inflammasome by caspase-1 adaptors ASC and Ipaf.
Nature 430: 213–218.
Mazaki Y, Hashimoto S, Tsujimura T, Morishige M, Hashimoto A, Aritake K, Yamada A, Nam JM,
Kiyonari H, Nakao K, et al. 2006. Neutrophil direction sensing and superoxide production linked by
the GTPase-activating protein GIT2. Nat Immunol 7: 724–731.
Miao EA, Leaf IA, Treuting PM, Mao DP, Dors M, Sarkar A, Warren SE, Wewers MD, Aderem A.
2010. Caspase-1-induced pyroptosis is an innate immune effector mechanism against intracellular
bacteria. Nat Immunol 11: 1136–1142.
Michallet MC, Meylan E, Ermolaeva MA, Vazquez J, Rebsamen M, Curran J, Poeck H, Bscheider M,
Hartmann G, Konig M, et al. 2008. TRADD protein is an essential component of the RIG-like
helicase antiviral pathway. Immunity 28: 651–661.
Micheau O, Tschopp J. 2003. Induction of TNF receptor I-mediated apoptosis via two sequential
signaling complexes. Cell 114: 181–190.
Micheau O, Lens S, Gaide O, Alevizopoulos K, Tschopp J. 2001. NF-κB signals induce the expression
of c-FLIP. Mol Cell Biol 21: 5299–5305.
Mukhopadhyay S, Plüddemann A, Hoe JC, Williams KJ, Varin A, Makepeace K, Aknin M, Bowdish
DME, Smale ST, Barclay AN, et al. 2010. Immune inhibitory ligand CD200 induction by TLRs and
NLRs limits macrophage activation to protect the host from meningococcal septicemia. Cell Host
Microbe 16: 236–247.
Muller WA. 2011. Mechanisms of leukocyte transendothelial migration. Annu Rev Pathol 28: 323–344.
Muzio M, Ni J, Feng P, Dixit VM. 1997. IRAK (Pelle) family member IRAK-2 and MyD88 as
proximal mediators of IL-1 signaling. Science 278: 1612–1615.
Newton K, Matsumoto ML, Wertz IE, Kirkpatrick DS, Lill JR, Tan J, Dugger D, Gordon N, Sidhu SS,
Fellouse FA, et al. 2008. Ubiquitin chain editing revealed by polyubiquitin linkage-specific
antibodies. Cell 134: 668–678.
Nishio M, Watanabe K, Sasaki J, Taya C, Takasuga S, Iizuka R, Balla T, Yamazaki M, Watanabe H,
Itoh R, et al. 2007. Control of cell polarity and motility by the PtdIns(3,4,5)P3 phosphatase SHIP1.
Nat Cell Biol 9: 36–44.
Park JH, Kim YG, McDonald C, Kanneganti TD, Hasegawa M, Body-Malapel M, Inohara N, Nunez G.
2007. RICK/RIP2 mediates innate immune responses induced through Nod1 and Nod2 but not
TLRs. J Immunol 178: 2380–2386.
Park EJ, Lee JH, Yu GY, He G, Ali SR, Holzer RG, Osterreicher CH, Takahashi H, Karin M. 2010.
Dietary and genetic obesity promote liver inflammation and tumorigenesis by enhancing IL-6 and
TNF expression. Cell 140: 197–208.
Peschon JJ, Torrance DS, Stocking KL, Glaccum MB, Otten C, Willis CR, Charrier K, Morrissey PJ,
Ware CB, Mohler KM. 1998. TNF receptor-deficient mice reveal divergent roles for p55 and p75 in
several models of inflammation. J Immunol 160: 943–952.
Pobezinskaya YL, Kim YS, Choksi S, Morgan MJ, Li T, Liu C, Liu Z. 2008. The function of TRADD
in signaling through tumor necrosis factor receptor 1 and TRIF-dependent Toll-like receptors. Nat
Immunol 9: 1047–1054.
Poeck H, Bscheider M, Gross O, Finger K, Roth S, Rebsamen M, Hannesschlager N, Schlee M,
Rothenfusser S, Barchet W, et al. 2010. Recognition of RNA virus by RIG-I results in activation of
CARD9 and inflammasome signaling for interleukin 1β production. Nat Immunol 11: 63–69.
Qu Y, Misaghi S, Izrael-Tomasevic A, Newton K, Gilmour LL, Lamkanfi M, Louie S, Kayagaki N, Liu
J, Komuves L, et al. 2012. Phosphorylation of NLRC4 is critical for inflammasome activation.
Nature 490: 539–542.
Rahighi S, Ikeda F, Kawasaki M, Akutsu M, Suzuki N, Kato R, Kensche T, Uejima T, Bloor S,
Komander D, et al. 2009. Specific recognition of linear ubiquitin chains by NEMO is important for
NF-κB activation. Cell 136: 1098–1109.
Reiley W, Zhang M, Wu X, Granger E, Sun SC. 2005. Regulation of the deubiquitinating enzyme
CYLD by IκB kinase γ-dependent phosphorylation. Mol Cell Biol 25: 3886–3895.
Rothlin CV, Ghosh S, Zuniga EI, Oldstone MB, Lemke G. 2007. TAM receptors are pleiotropic
inhibitors of the innate immune response. Cell 131: 1124–1136.
Sato S, Sanjo H, Takeda K, Ninomiya-Tsuji J, Yamamoto M, Kawai T, Matsumoto K, Takeuchi O,
Akira S. 2005. Essential function for the kinase TAK1 in innate and adaptive immune responses.
Nat Immunol 6: 1087–1095.
Satoh T, Kato H, Kumagai Y, Yoneyama M, Sato S, Matsushita K, Tsujimura T, Fujita T, Akira S,
Takeuchi O. 2010. LGP2 is a positive regulator of RIG-I- and MDA5-mediated antiviral responses.
Proc Natl Acad Sci 107: 1512–1517.
Shembade N, Ma A, Harhaj EW. 2010. Inhibition of NF-κB signaling by A20 through disruption of
ubiquitin enzyme complexes. Science 327: 1135–1139.
Shi Y, Zhang J, Mullin M, Dong B, Alberts AS, Siminovitch KA. 2009. The mDial formin is required
for neutrophil polarization, migration, and activation of the LARG/RhoA/ROCK signaling axis
during chemotaxis. J Immunol 182: 3837–3845.
Shim JH, Xiao C, Paschal AE, Bailey ST, Rao P, Hayden MS, Lee KY, Bussey C, Steckel M, Tanaka
N, et al. 2005. TAK1, but not TAB1 or TAB2, plays an essential role in multiple signaling
pathways in vivo. Genes Dev 19: 2668–2681.
Shisler JL, Jin XL. 2004. The vaccinia virus K1L gene product inhibits host NF-κB activation by
preventing IκBα degradation. J Virol 78: 3553–3560.
Skaug B, Jiang X, Chen ZJ. 2009. The role of ubiquitin in NF-κB regulatory pathways. Annu Rev
Biochem 78: 769–796.
Suire S, Condliffe AM, Ferguson GJ, Ellson CD, Guillou H, Davidson K, Welch H, Coadwell J, Turner
M, Chilvers ER, et al. 2006. Gβγs and the Ras binding domain of p110γ are both important
regulators of PI(3)Kγ signalling in neutrophils. Nat Cell Biol 8: 1303–1309.
Sun CX, Downey GP, Zhu F, Koh AL, Thang H, Glogauer M. 2004a. Rac1 is the small GTPase
responsible for regulating the neutrophil chemotaxis compass. Blood 104: 3758–3765.
Sun L, Deng L, Ea CK, Xia ZP, Chen ZJ. 2004b. The TRAF6 ubiquitin ligase and TAK1 kinase
mediate IKK activation by BCL10 and MALT1 in T lymphocytes. Mol Cell 14: 289–301.
Sun L, Wang H, Wang Z, He S, Chen S, Liao D, Wang L, Yan J, Liu W, Lei X, et al. 2012. Mixed
lineage kinase domain-like protein mediates necrosis signaling downstream of RIP3 kinase. Cell
148: 213–227.
Sun L, Wu J, Du F, Chen X, Chen ZJ. 2013. Cyclic GMP-AMP synthase is a cytosolic DNA sensor
that activates the type I interferon pathway. Science 339: 786–791.
Suzuki K, Verma IM. 2008. Phosphorylation of SNAP-23 by IκB kinase 2 regulates mast cell
degranulation. Cell 134: 485–495.
Takahashi H, Ogata H, Nishigaki R, Broide DH, Karin M. 2010. Tobacco smoke promotes lung
tumorigenesis by triggering IKKβ- and JNK1-dependent inflammation. Cancer Cell 17: 89–97.
Takeuchi O, Akira S. 2010. Pattern recognition receptors and inflammation. Cell 140: 805–820.
Taylor SL, Frias-Staheli N, Garcia-Sastre A, Schmaljohn CS. 2009. Hantaan virus nucleocapsid protein
binds to importin α proteins and inhibits tumor necrosis factor α-induced activation of nuclear
factor κB. J Virol 83: 1271–1279.
Tokunaga F, Sakata S, Saeki Y, Satomi Y, Kirisako T, Kamei K, Nakagawa T, Kato M, Murata S,
Yamaoka S, et al. 2009. Involvement of linear polyubiquitylation of NEMO in NF-κB activation.
Nat Cell Biol 11: 123–132.
Tokunaga F, Nakagawa T, Nakahara M, Saeki Y, Taniguchi M, Sakata S, Tanaka K, Nakano H, Iwai
K. 2011. SHARPIN is a component of the NF-κB-activating linear ubiquitin chain assembly
complex. Nature 471: 633–636.
Travassos LH, Carneiro LA, Ramjeet M, Hussey S, Kim YG, Magalhaes JG, Yuan L, Soares F, Chea
E, Le Bourhis L, et al. 2010. Nod1 and Nod2 direct autophagy by recruiting ATG16L1 to the
plasma membrane at the site of bacterial entry. Nat Immunol 11: 55–62.
Tseng PH, Matsuzawa A, Zhang W, Mino T, Vignali DA, Karin M. 2010. Different modes of
ubiquitination of the adaptor TRAF3 selectively activate the expression of type I interferons and
proinflammatory cytokines. Nat Immunol 11: 70–75.
Utomo A, Cullere X, Glogauer M, Swat W, Mayadas TN. 2006. Vav proteins in neutrophils are
required for FcγR-mediated signaling to Rac GTPases and nicotinamide adenine dinucleotide
phosphate oxidase component p40(phox). J Immunol 177: 6388–6397.
Varfolomeev E, Goncharov T, Fedorova AV, Dynek JN, Zobel K, Deshayes K, Fairbrother WJ, Vucic
D. 2008. c-IAP1 and c-IAP2 are critical mediators of tumor necrosis factor α (TNFα)-induced NF-
κB activation. J Biol Chem 283: 24295–24299.
Venkataraman T, Valdes M, Elsby R, Kakuta S, Caceres G, Saijo S, Iwakura Y, Barber GN. 2007. Loss
of DExD/H box RNA helicase LGP2 manifests disparate antiviral responses. J Immunol 178: 6444–
6455.
Vereecke L, Beyaert R, van Loo G. 2011. Genetic relationships between A20/TNFAIP3, chronic
inflammation, and autoimmune disease. Biochem Soc Trans 39: 1086–1091.
Vig M, DeHaven WI, Bird GS, Billingsley JM, Wang H, Rao PE, Hutchings AB, Jouvin MH, Putney
JW, Kinet JP. 2008. Defective mast cell effector functions in mice lacking the CRACM1 pore
subunit of store-operated calcium release-activated calcium channels. Nat Immunol 9: 89–96.
Wan Y, Xiao H, Affolter J, Kim TW, Bulek K, Chaudhuri S, Carlson D, Hamilton T, Mazumder B,
Stark GR, et al. 2009. Interleukin-1 receptor-associated kinase 2 is critical for lipopolysaccharide-
mediated post-transcriptional control. J Biol Chem 284: 10367–10375.
Wang J, Basagoudanavar SH, Wang X, Hopewell E, Albrecht R, Garcia-Sastre A, Balachandran S, Beg
AA. 2010a. NF-κB RelA subunit is crucial for early IFN-β expression and resistance to RNA virus
replication. J Immunol 185: 1720–1729.
Wang Y, Ludwig J, Schuberth C, Goldeck M, Schlee M, Li H, Juranek S, Sheng G, Micura R, Tuschl
T, et al. 2010b. Structural and functional insights into 5′-ppp RNA pattern recognition by the innate
immune receptor RIG-I. Nat Struct Mol Biol 17: 781–787.
Welch HC, Coadwell WJ, Ellson CD, Ferguson GJ, Andrews SR, Erdjument-Bromage H, Tempst P,
Hawkins PT, Stephens LR. 2002. P-Rex1, a PtdIns(3,4,5)P3- and Gβγ-regulated guanine-nucleotide
exchange factor for Rac. Cell 108: 809–821.
Wong WW, Gentle IE, Nachbur U, Anderton H, Vaux DL, Silke J. 2010. RIPK1 is not essential for
TNFR1-induced activation of NF-κB. Cell Death Differ 17: 482–487.
Wu J, Sun L, Chen X, Du F, Shi H, Chen C, Chen ZJ. 2013a. Cyclic GMP-AMP is an endogenous
second messenger in innate immune signaling by cytosolic DNA. Science 339: 826–830.
Wu J, Huang Z, Ren J, Zhang Z, He P, Li Y, Ma J, Chen W, Zhang Y, Zhou X, et al. 2013b. Mlkl
knockout mice demonstrate the indispensable role of Mlkl in necroptosis. Cell Res 23: 994–1006.
Xia ZP, Sun L, Chen X, Pineda G, Jiang X, Adhikari A, Zeng W, Chen ZJ. 2009. Direct activation of
protein kinases by unanchored polyubiquitin chains. Nature 461: 114–119.
Xu M, Skaug B, Zeng W, Chen ZJ. 2009. A ubiquitin replacement strategy in human cells reveals
distinct mechanisms of IKK activation by TNFα and IL-1β. Mol Cell 36: 302–314.
Yago T, Shao B, Miner JJ, Yao L, Klopocki AG, Maeda K, Coggeshall KM, McEver RP. 2010. E-
selectin engages PSGL-1 and CD44 through a common signaling pathway to induce integrin αLβ2-
mediated slow leukocyte rolling. Blood 116: 485–494.
Yamamoto M, Okamoto T, Takeda K, Sato S, Sanjo H, Uematsu S, Saitoh T, Yamamoto N, Sakurai H,
Ishii KJ, et al. 2006. Key function for the Ubc13 E2 ubiquitin-conjugating enzyme in immune
receptor signaling. Nat Immunol 7: 962–970.
Yang Y, Yin C, Pandey A, Abbott D, Sassetti C, Kelliher MA. 2007. NOD2 pathway activation by
MDP or Mycobacterium tuberculosis infection involves the stable polyubiquitination of Rip2. J
Biol Chem 282: 36223–36229.
Yang YJ, Chen W, Carrigan SO, Chen WM, Roth K, Akiyama T, Inoue J, Marshall JS, Berman JN, Lin
TJ. 2008. TRAF6 specifically contributes to FcεRI-mediated cytokine production but not mast cell
degranulation. J Biol Chem 283: 32110–32118.
Yoshida R, Takaesu G, Yoshida H, Okamoto F, Yoshioka T, Choi Y, Akira S, Kawai T, Yoshimura A,
Kobayashi T. 2008. TRAF6 and MEKK1 play a pivotal role in the RIG-I-like helicase antiviral
pathway. J Biol Chem 283: 36211–36220.
Zarbock A, Abram CL, Hundt M, Altman A, Lowell CA, Ley K. 2008. PSGL-1 engagement by E-
selectin signals through Src kinase Fgr and ITAM adapters DAP12 and FcRγ to induce slow
leukocyte rolling. J Exp Med 205: 2339–2347.
Zeng W, Sun L, Jiang X, Chen X, Hou F, Adhikari A, Xu M, Chen ZJ. 2010. Reconstitution of the
RIG-I pathway reveals a signaling role of unanchored polyubiquitin chains in innate immunity. Cell
141: 315–330.
Zhang J, Guo J, Dzhagalov I, He YW. 2005. An essential function for the calcium-promoted Ras
inactivator in Fcγ receptor-mediated phagocytosis. Nat Immunol 6: 911–919.
Zhang DW, Shao J, Lin J, Zhang N, Lu BJ, Lin SC, Dong MQ, Han J. 2009. RIP3, an energy
metabolism regulator that switches TNF-induced cell death from apoptosis to necrosis. Science 325:
332–336.
Zhao T, Yang L, Sun Q, Arguello M, Ballard DW, Hiscott J, Lin R. 2007. The NEMO adaptor bridges
the nuclear factor-κB and interferon regulatory factor signaling pathways. Nat Immunol 8: 592–600.
1DCs are very heterogeneous. pDCs acquire DC morphology and secrete large amounts of IFN during
virus infections. They can be distinguished from other DC subsets by their cell surface markers. The
myeloid DCs referenced were derived in vitro from bone marrow cells with granulocyte/macrophage
colony-stimulating factor (GM-CSF).
2Autophagy is the process by which cytoplasmic components, including organelles and invading
bacteria, are sequestered inside double-membrane vesicles and then delivered to the lysosome for
degradation.
3Chemokinesis refers to random cell migration, whereas chemotaxis is directed cell migration along a
chemical gradient.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a006049
CHAPTER 16
Doreen Cantrell
College of Life Sciences, Wellcome Trust Biocentre, University of Dundee,
Dundee DD1 5EH, Scotland, United Kingdom
Correspondence: d.a.cantrell@dundee.ac.uk
SUMMARY
The fate of T and B lymphocytes, the key cells that direct the adaptive
immune response, is regulated by a diverse network of signal
transduction pathways. The T- and B-cell antigen receptors are coupled
to intracellular tyrosine kinases and adaptor molecules to control the
metabolism of inositol phospholipids and calcium release. The
production of inositol polyphosphates and lipid second messengers
directs the activity of downstream guanine-nucleotide-binding proteins
and protein and lipid kinases/phosphatases that control lymphocyte
transcriptional and metabolic programs. Lymphocyte activation is
modulated by costimulatory molecules and cytokines that elicit
intracellular signaling that is integrated with the antigen-receptor-
controlled pathways.
Outline
1 Introduction
2 Antigen-receptor structure and function
3 Immunoreceptor tyrosine-based activation motifs
4 Adaptor molecules for antigen receptors
5 Calcium and diacylglycerol signaling
6 Downstream from calcium signaling in lymphocytes
7 Diacylglycerol signaling in lymphocytes
8 PKC and lymphocytes
9 Ras signaling and lymphocytes
10 Costimulatory molecules, cytokines, and lymphocyte activation
11 Cytokine signaling in lymphocytes
12 PI3K-mediated signaling in lymphocytes
13 Inhibitory signals and lymphocyte activation
14 Concluding remarks
References
1 INTRODUCTION
The adaptive immune response is directed by B and T lymphocytes. These
cells express specific receptors that recognize pathogen-derived antigens: the
B-cell antigen receptor (BCR) and the T-cell antigen receptor (TCR),
respectively. B lymphocytes have two principal roles: to produce and secrete
specific antibodies/immunoglobulins, and to function as antigen-presenting
cells (APCs). T cells have multiple roles in adaptive immune responses. In
this context, peripheral T cells can be subdivided on the basis of whether they
express CD8 or CD4, receptors that recognize class I and class II major
histocompatibility complex (MHC) molecules, respectively. CD8+ T cells
differentiate to cytolytic effectors that directly kill virus- or bacteria-infected
cells. CD4+ T cells are referred to as “helper” T cells because they produce
regulatory cytokines and chemokines that mediate autocrine or paracrine
control of T-cell differentiation and/or regulate the differentiation of B cells
and/or direct the activity of macrophages and neutrophils (O’Shea and Paul
2010). At least five major subpopulations of mature CD4+ cells exist with
distinct functions that are tailored to deal with different pathogens. Th1 cells,
characterized by interferon (IFN)γ production; Th2 cells, characterized by
interleukin 4 (IL4) and IL13 production; Th17 cells, which produce the
proinflammatory cytokines IL17 and IL22; regulatory T (Treg) cells that
function to restrain autoimmunity and strong inflammatory responses; and
follicular helper T (Tfh) cells, a class of effector CD4+ T cells that regulate
the development of antigen-specific B-cell immunity.
The paradigm of the adaptive immune response is that a primary response
to an antigen causes clonal expansion of antigen-reactive T or B cells and
produces a large number of effector lymphocytes that cause clearance of the
pathogen. Once the pathogen is cleared there is a contraction phase of the
immune response characterized by loss of effector lymphocytes and the
emergence of long-lived memory cells capable of mounting rapid secondary
responses to reinfection with the original pathogen.
The proliferation and differentiation of mature lymphocytes in adaptive
immune responses are directed by antigen receptors, costimulatory
molecules, adhesion molecules, cytokines, and chemokines. These extrinsic
stimuli are coupled to a diverse network of signal transduction pathways that
control the transcriptional and metabolic programs that determine lymphocyte
function. At the core of lymphocyte signal transduction is the regulated
metabolism of inositol phospholipids and the resultant production of inositol
polyphosphates and lipids such as polyunsaturated diacylglycerols (DAGs).
These second messengers direct the activity of protein and lipid kinases and
guanine-nucleotide-binding proteins that control lymphocyte proliferation,
differentiation, and effector function. Below, I outline both the unique and
the conserved aspects of signaling in lymphocytes, focusing on signaling
pathways controlled by antigen receptors and how these responses are
subsequently shaped and modulated by cytokines and chemokines.
3 IMMUNORECEPTOR TYROSINE-BASED
ACTIVATION MOTIFS
The antigen-receptor subunits that mediate signal transduction are the
invariant chains CD3γ, δ, ε, ζ in T cells, Igα and Igβ in B lymphocytes, and
the FcRγ chain in mast cells (see below). These signaling subunits have no
intrinsic signaling capacity, but all contain a YxxL/I-X6–8-YxxL/I motif
referred to as an immunoreceptor tyrosine-based activation motif (ITAM)
(Abram and Lowell 2007; Love and Hayes 2010). The CD3γ, δ, and ε
subunits each contain a single ITAM, and there are three ITAMs in the CD3ζ
chain. The minimal TCR complex thus has 10 ITAMs. These couple the TCR
to intracellular tyrosine kinases (see below). ITAM motifs are a defining
feature of antigen-receptor complexes. Igα and Igβ, the signaling subunits of
the BCR, both have a single ITAM.
ITAM motifs are not restricted to the TCR and BCR. For example, mast
cells comprise an important group of lymphocytes whose fate is determined
by antigen-specific immunoglobulin. These cells respond to antigen because
they express a high-affinity receptor for IgE. This receptor, termed FcεR1,
binds to the immunoglobulin IgE with high affinity. When FcεR1-IgE
complexes are cross-linked by polyvalent antigen they can trigger mast cell
degranulation and the release of cytokines and allergic mediators. The FcεR1
is assembled from three subunits: the α subunit that binds to the Fc region of
IgE, a β subunit that provides important accessory signaling, and the FcRγ
chain, which is a signaling subunit that contains a single ITAM (Beaven and
Metzger 1993; Abram and Lowell 2007; p. 125 [Samelson 2011]).
TCR/BCR/FcεR1 signaling is initiated by the tyrosine phosphorylation of
ITAMs by Src-family tyrosine kinases such as Lck and Fyn in T cells, Lyn in
B cells, and Fyn in mast cells (Salmond et al. 2009). When both tyrosine
residues are phosphorylated, the ITAM forms a high-affinity binding site for
Syk-family tyrosine kinases; generally in T cells this is Zap-70 (Wang et al.
2010), whereas in B cells and mast cells Syk is recruited (Chu et al. 1998).
Zap-70 and Syk contain tandem SH2 domains that bind with high affinity to
the doubly phosphorylated ITAM (Chu et al. 1998). The activation of Zap-70
or Syk is initiated by binding to phosphorylated ITAMs. This is proposed to
release Syk/Zap-70 from an autoinhibited conformation and expose
regulatory tyrosine residues for phosphorylation by Src-family kinases (Au-
Yeung et al. 2009). The phosphorylation of tyrosine residues in the activation
loop in the Zap-70/Syk catalytic domain, as well as two residues in the
adjacent linker region, then further stimulates their catalytic activity.
Antigen-receptor control of Syk-family tyrosine kinases is fundamental for
lymphocyte activation and underpins the ability of antigen receptors to
transduce signals from pathogen-derived antigens to the interior of
lymphocytes (Mocsai et al. 2010; Wang et al. 2010).
How the Src-family kinases such as Lck are regulated is central to
antigen-receptor signal transduction (Salmond et al. 2009). The activity of
Lck is regulated by phosphorylation and dephosphorylation of a carboxy-
terminal tyrosine (Y505) by the ubiquitously expressed kinase carboxy-
terminal Src kinase (CSK), as well as autophosphorylation of the activation
loop tyrosine residue, Y394. Phosphorylated Y505 forms an intramolecular
binding site for the Lck SH2 domain, thereby locking the kinase into an
autoinhibited state. The key to initiating the activation of Lck and its relatives
is to dephosphorylate the carboxy-terminal tyrosine and relieve autoinhibition
of the kinase. This is mediated by transmembrane-receptor-like tyrosine
phosphatases, such as CD45 and CD148 (Hermiston et al. 2009; Zikherman
et al. 2010). Hence in T cells, the Lck activation threshold is set by the
balanced activity of the kinase-phosphatase pair CSK, which phosphorylates
Y505, and CD45, which dephosphorylates this residue (Zikherman et al.
2010).
It is frequently assumed that triggering antigen receptors stimulates Src
kinase family activity, and antigen receptors are often depicted as molecular
switches that are either on or off. In reality, antigen receptors are always
signaling and it is the intensity of the signal that changes. The assembly of
antigen receptors at the plasma membrane is thus proposed to mediate low-
level signaling and the engagement with high-affinity ligands (antigen or
antigen–MHC) increases the intensity. Indeed Src-family kinases such as Lck
are constitutively active before antigen-receptor engagement and cause low-
level ITAM phosphorylation (Nika et al. 2010). The levels of ITAM
phosphorylation are limited by tyrosine phosphatases, and the increases in
ITAM phosphorylation that follow antigen-receptor engagement probably
result from spatial constraints on the ITAM-phosphatase interaction (van der
Merwe and Dushek 2011).
How are these spatial constraints regulated to explain how ligand
occupancy triggers TCR signaling? Surprisingly, we do not know, although
there is no shortage of theories. Current models range from the ligand-
induced conformational change to the idea that the TCR is a mechanosensor
that converts the mechanical energy generated by antigen binding into a
biochemical signal (Kim et al. 2009). One other idea well supported by
experimental data is that binding of the TCR to peptide-MHC complexes on
the surface ofAPCs causes spatial segregation of TCR complexes (which
have small ectodomains) away from receptor tyrosine phosphatases such as
CD45 and CD148 (which have very large ectodomains). This might locally
perturb the kinase–phosphatase balance sufficiently to favor ITAM
phosphorylation and Zap-70 recruitment (van der Merwe and Davis 2003;
van der Merwe and Dushek 2011). Note that the MHC-binding coreceptors
CD4 and CD8 are also thought to play a role in perturbing the kinase-
phosphatase balance in localized areas of the T-cell membrane. CD4 and
CD8 can thus promote TCR signaling by stabilizing interactions between the
TCR and peptide-MHC ligands. However, the cytoplasmic domains of CD4
and CD8 constitutively bind Lck and hence facilitate the recruitment of this
kinase to ligand-engaged TCR complexes (Artyomov et al. 2010).
What about the BCR and FcεR1? In quiescent B cells, the BCR may exist
in an oligomeric autoinhibited state, and ligand occupancy could drive the
dissociation of these oligomers into monomers that interact more effectively
with downstream tyrosine kinases (Yang and Reth 2010a,b). For the FcεR1,
the opposite is probably the case. This receptor binds IgE but is only
effectively triggered when antigen oligomerizes the receptor (Beaven and
Metzger 1993).
10 COSTIMULATORY MOLECULES,
CYTOKINES, AND LYMPHOCYTE
ACTIVATION
Lymphocyte responses both prior and subsequent to antigen-receptor
engagement are modulated by multiple costimulatory and coinhibitory
receptors. Signaling via Toll-like receptors (TLRs) is also a major factor
influencing the fate of lymphocytes during an immune response. Because T
and B lymphocytes respond to antigens presented to them by APCs,
lymphocyte activation can be regulated by the adhesion molecules and
costimulatory molecules expressed by the APC. Note also that many of the
cytokines that control lymphocyte fate are produced in response to TLR-
mediated activation of dendritic cells and macrophages (Ch. 15 [Newton and
Dixit 2012]). Hence, the nature of the pathogen challenge to the innate
immune system, and the resultant cytokine milieu modulate the adaptive
immune response.
For T cells, key coreceptor molecules include the MHC receptors CD4
and CD8, and proteins such as CD28 (a positive coregulator) and CTLA4 and
PD-1 (negative coregulators) (Artyomov et al. 2010; Francisco et al. 2010;
Bour-Jordan et al. 2011; Walker and Sansom 2011). In B cells, molecules
such as CD19 and the CD21 receptor for complement component C3d are
essential (Carter and Fearon 1992; Depoil et al. 2008; Elgueta et al. 2009;
Mackay et al. 2010) as are the TNF receptor family members CD40 and
receptor for B-cell-activating factor (BAFFR) (Watts 2005; Elgueta et al.
2009; Karin and Gallagher 2009).
A full review of lymphocyte regulation by costimulatory factors is beyond
our scope here but there are some general themes. Costimulatory molecules
frequently work as adaptors to recruit signaling molecules to the plasma
membrane and hence amplify antigen-receptor-mediated signaling. For
example, CD4 and CD8 in T cells recruit Lck to the plasma membrane.
Similarly, CD28 in T cells and CD19 in B cells both have cytoplasmic
domains that can be tyrosine phosphorylated and thus can act as docking sites
for SH2-domain-containing adaptors and enzymes. The CD19 cytoplasmic
tail contains nine tyrosine residues with the potential to be phosphorylated
and interact with signaling molecules including lipid kinases, Vav-family
GEFs, and adaptor proteins such as Grb2. Other important examples of
molecules that recruit key adaptor molecules to the plasma membrane are the
lymphocytic activation molecule (SLAM) family of receptors and associated
intracellular adaptors of the SLAM-associated protein (SAP) family (Veillette
2010).
The engagement of CD40 by its ligand (CD40L) leads to signals via
adaptor proteins known as TNFR-associated factors (TRAFs), which activate
signaling pathways, including MAPKs and NF-κB (p. 121 [Lim and Staudt
2012]).
The plethora of costimulatory molecules that can contribute to
lymphocyte activation can be confusing, particularly because all seem to
activate similar signal transduction pathways. The key message is that these
receptors function at different times and in different contexts. For example,
CD28 binds to the B7 family members CD80 and CD86, which are mainly
expressed on APCs responding to TLR signaling. The ligand for CD40 is
produced transiently by antigen-activated T cells and plays a key role in
promoting specific T cell “help” to B cells by ensuring integration of signals
between CD40-expressing B cells and antigen-primed T cells. In contrast,
BAFF is mainly produced by neutrophils, monocytes, and macrophages and
hence allows cross talk between B cells and these cells of the innate immune
system.
11 CYTOKINE SIGNALING IN
LYMPHOCYTES
Cytokines that signal via the Janus tyrosine kinases (JAKs) (p. 117 [Harrison
2012]), such as the γc family of cytokines, IFNs, and cytokines such as IL12
and IL23, are particularly important to the adaptive immune system
(Rochman et al. 2009). For example, CD4-expressing αβ T cells differentiate
during immune responses to produce distinct effector subpopulations
(O’Shea and Paul 2010) and the specification of these CD4+ T-cell subsets is
controlled by cytokines that direct the combinatorial action of multiple
chromatin regulators and key lineage-specifying transcription factors. For
example, IL12 drives Th1 T-cell differentiation and IL6, IL21, and IL23
drive Th17 cell differentiation. Moreover, cytokines have pleotropic roles.
IL2 is important for the differentiation of antigen-primed CD8+ T cells to
effector cytotoxic T cells (CTLs) but is also required for optimal Th1 T-cell
differentiation and for the development of Treg cells.
One striking feature of lymphocyte biology is that the ability of cells to
respond to cytokines (i.e., to express particular cytokine receptors) can be
shaped by antigen-receptor triggering. Cytokine production by cells of the
immune system is, in turn, controlled by triggering of antigen receptors in T
and B cells or by receptors of the innate immune system. A prototypical
example is IL2, which is only produced by antigen-receptor-activated T cells
and B cells or pathogen-triggered dendritic cells. Moreover, expression of the
IL2 receptor (IL2R) is tightly controlled by immune activation. The ILR2
receptor complex consists of a γc, a β subunit (CD122), and an α subunit
(CD25). The expression of CD25 is rate limiting as it determines the ability
of the receptor to bind IL2 with high affinity. Importantly, CD25 is not
expressed on naïve CD4 and CD8 T cells but only on activated T cells. In
addition, the expression of CD25 is transient and its sustained expression
requires constant immune stimulation. IL2 responsiveness is thus tightly
linked to antigen-receptor triggering to ensure the tight control of T cells by
IL2. IL12 receptors are similar: these are only expressed on activated T cells.
Furthermore, IL12 receptor expression needs to be sustained by IL2 and there
is tight control of IL12 secretion by pathogen-activated dendritic cells and
macrophages. Such dynamic regulation of cytokine and cytokine-receptor
expression during immune activation ensures the immune specificity of
cytokine action (i.e., only lymphocytes that have been primed by antigen-
receptor triggering can respond to IL12). Note the production of cytokines is
also limited to either pathogen-activated innate immune cells or antigen-
activated lymphocytes (Fig. 2).
Figure 2. Signaling by interleukin (IL) receptors. Many cytokines signal via receptors linked to Janus
tyrosine kinases (JAKs), which regulate the SH2-domain-containing transcription factors STATs. The
different ILs produced by different cell types activate receptors coupled to different combinations of
JAKs and STATs.
12 PI3K-MEDIATED SIGNALING IN
LYMPHOCYTES
PI3K signaling is important for lymphocyte activation and integrates multiple
receptor inputs. For example, in naïve T cells, low basal levels of PIP3 are
maintained by IL7 signaling; these increase strikingly in response to
triggering of the antigen-receptor complex and are then sustained by stimuli
from costimulatory molecules such as CD28. Cytokines such as IL2 and IL15
can then further sustain intracellular concentrations of PIP3. Similarly, in B
cells, cytokines such as BAFF and low-level signaling by non-antigen-
engaged BCRs maintain a low level of PIP3 (Srinivasan et al. 2009). The
levels of PIP3 increase following BCR activation, and costimulatory
molecules such as CD19 and cytokines such as IL4 can also sustain levels of
this lipid.
Antigen receptor and cytokines control PIP3 metabolism in lymphocytes
via class I PI3Ks, which typically exist in a complex comprising a p110
catalytic subunit and an 85-kDa SH2-domain-containing regulatory/adaptor
subunit. Four p110 isoforms exist (α, β, γ, and δ) and two p85 subunits (α and
β) exist. These different isoforms function in distinct pathways in
lymphocytes, and expression of p110δ is restricted to hematopoietic cells.
p110δ produces the PIP3 that is generated in response to many antigen
receptors and cytokines, whereas p110γ, which heterodimerizes with the
p101 regulatory subunit rather than a p85-type subunit, is involved in
chemokine receptor signaling (Okkenhaug and Fruman 2010).
The production of PIP3 requires recruitment of PI3K to the plasma
membrane. There are two possible mechanisms: binding of the SH2 domain
of p85 to phosphorylated tyrosine residues in receptor cytoplasmic domains
or membrane-localized adaptors; and direct recruitment of p110 by Ras. In
the case of the BCR, CD19 recruits PI3K to the plasma membrane via
binding of p85 to its tyrosine-phosphorylated cytoplasmic domain. Tyrosine-
phosphorylated cytokine receptors similarly recruit PI3K by binding p85.
Surprisingly, how TCR and CD28 signaling induces PIP3 accumulation is not
known, but direct recruitment to tyrosine-phosphorylated CD28 does not
occur, and it is more likely that adaptors such as LAT or SLP76 are
important.
PIP3 binds to pleckstrin homology (PH) domains in other signaling
proteins to control their activity and subcellular localization. In lymphocytes,
these include Tec-family tyrosine kinases such as Itk and Btk, GEFs for Rho
family GTPases, and the kinases PDK1 and Akt (also known as PKB) (p. 87
[Hemmings and Restuccia 2012]). Akt is activated by PDK1-mediated
phosphorylation of T308 within its catalytic domain. This is PIP3 dependent
probably because the binding of PIP3 to the Akt PH domain causes a
conformational change that allows PDK1 to phosphorylate T308. PDK1 also
has a PIP3-binding PH domain, but this promotes translocation of the enzyme
to the plasma membrane (where it can colocalize with Akt) rather than
enzyme activation (Finlay and Cantrell 2011).
Once activated, Akt phosphorylates a number of critical signaling
molecules. For example, it phosphorylates and inactivates the Rheb GAP
TSC2, causing accumulation of Rheb-GTP complexes, which play a role in
activating the mTORC1 complex (mammalian target of rapamycin complex
1) (p. 91 [Laplante and Sabatini 2012]). Akt also phosphorylates the
transcription factors Foxo1/3 and Fox4A. These Foxo family transcription
factors are nuclear and active in quiescent cells but, when phosphorylated,
they exit the nucleus and form a complex with 14-3-3 proteins in the cytosol,
which terminates their transcriptional activity.
Akt is fundamentally important in many cells because it controls nutrient
uptake and cellular metabolism. In particular, activated lymphocytes up-
regulate glucose, amino acid and iron uptake, and switch their metabolism to
glycolysis (see Ch. 7 [Ward and Thompson 2012]). This increases cellular
energy production and nutrient uptake to support the increased biosynthetic
demands of rapid cell proliferation. Note, however, that it is difficult to
ascribe a universal function for Akt that holds for all lymphocyte
subpopulations. For example, Akt is important for metabolism and cell
survival in peripheral B lymphocytes (Srinivasan et al. 2009) and in T
lymphocyte progenitors in the thymus, but is not essential for metabolism or
for the survival of peripheral or effector cytotoxic T cells (Finlay and Cantrell
2011). Moreover, the Akt/Foxo pathway has a critical role controlling
expression of the recombinase genes responsible for antigen-receptor
diversity in B cells (Kuo and Schlissel 2009) but there is no evidence for such
a role in T cells. The molecular basis for these differences is not understood
but probably reflects redundancies with other kinases that have similar
substrate specificities (e.g., SGK1).
Akt/Foxo signaling is also uniquely linked to the regulation of the
expression of key cytokine and chemokine receptors and adhesion molecules
in lymphocytes (Hedrick 2009; Lorenz 2009; Macintyre et al. 2011). Hence,
when Akt is inactive in quiescent lymphocytes, nonphosphorylated Foxo1,
Foxo3, Foxo3A, and Foxo4 are found in the nucleus, where they drive
transcription of genes encoding the receptor for IL7, an essential homeostatic
cytokine for lymphocytes. Moreover, Foxo transcription factors also drive
expression of the transcription factor KLF2; this directly regulates
transcription of adhesion molecules and chemokine receptors that together
control lymphocyte entry and egress from secondary lymphoid tissues and
lymphocyte positioning in lymphoid tissue. The activation of Akt thus causes
lymphocytes to change their trafficking program around the body. Akt
activation also changes the cytokine-receptor profile of T cells and hence the
ability of cytokines to determine T-cell fate.
In many cells, a key role for Akt is to control the activity of the
mammalian target of rapamycin complex 1 (mTORC1) (p. 91 [Laplante and
Sabatini 2012]). Rapamycin is a powerful immunosuppressant that is used in
the clinic to prevent rejection of organ transplants. mTORC1 coordinates
inputs from nutrients and antigen and cytokine receptors to control T-cell
differentiation (Powell and Delgoffe 2010). The molecular mechanisms used
by mTORC1 to control T-cell differentiation are not fully understood; neither
are the signaling processes that activate mTORC1. There is, however,
evidence that mTORC1 controls expression of genes encoding effector
cytokines and cytolytic molecules. Moreover, mTORC1 directs the tissue-
homing properties of T cells by regulating the expression of chemokine and
adhesion receptors (Sinclair et al. 2008).
14 CONCLUDING REMARKS
In lymphocytes, signal inputs generated by specific pathogens regulate the
activity of evolutionarily conserved signaling pathways. Antigen receptors
direct the immune response but lymphocyte signaling is also controlled by
cytokines and chemokines that are not antigen specific. These antigen-
specific and -nonspecific elements of lymphocyte signal transduction are
tightly coupled because antigen-receptor signaling controls the repertoire of
cytokine and chemokine receptors and adhesion molecules expressed by
lymphocytes. Antigen receptors also direct lymphocyte trafficking between
the blood, peripheral tissues, and secondary lymphoid organs and hence
control the cytokine milieu available to these cells. This coordination of
antigen receptor and cytokine signaling ensures the immune specificity of
lymphocyte activation and is fundamental for adaptive immune responses.
REFERENCES
*Reference is in this book.
Abram CL, Lowell CA. 2007. The expanding role for ITAM-based signaling pathways in immune
cells. Sci STKE 2007: re2.
Alarcón B, Swamy M, van Santen HM, Schamel WW. 2006. T-cell antigen-receptor stoichiometry:
Pre-clustering for sensitivity. EMBO Rep 7: 490–495.
Alvarez-Errico D, Lessmann E, Rivera J. 2009. Adapters in the organization of mast cell signaling.
Immunol Rev 232: 195–217.
Artyomov MN, Lis M, Devadas S, Davis MM, Chakraborty AK. 2010. CD4 and CD8 binding to MHC
molecules primarily acts to enhance Lck delivery. Proc Natl Acad Sci 107: 16916–16921.
Au-Yeung BB, Deindl S, Hsu LY, Palacios EH, Levin SE, Kuriyan J, Weiss A. 2009. The structure,
regulation, and function of ZAP-70. Immunol Rev 228: 41–57.
Beaven MA, Metzger H. 1993. Signal transduction by Fc receptors: The FcεRI case. Immunol Today
14: 222–226.
Blonska M, Lin X. 2009. CARMA1-mediated NF-κB and JNK activation in lymphocytes. Immunol
Rev 228: 199–211.
* Bootman MD. 2012. Calcium signaling. Cold Spring Harb Perspect Biol 4: a011171.
Bour-Jordan H, Esensten JH, Martinez-Llordella M, Penaranda C, Stumpf M, Bluestone JA. 2011.
Intrinsic and extrinsic control of peripheral T-cell tolerance by costimulatory molecules of the
CD28/B7 family. Immunol Rev 241: 180–205.
Cantrell DA. 2003. GTPases and T cell activation. Immunol Rev 192: 122–130.
Cao H, Crocker PR. 2011. Evolution of CD33-related siglecs: Regulating host immune functions and
escaping pathogen exploitation? Immunology 132: 18–26.
Carter RH, Fearon DT. 1992. CD19: Lowering the threshold for antigen receptor stimulation of B
lymphocytes. Science 256: 105–107.
Chakraborty AK, Das J, Zikherman J, Yang M, Govern CC, Ho M, Weiss A, Roose J. 2009. Molecular
origin and functional consequences of digital signaling and hysteresis during Ras activation in
lymphocytes. Sci Signal 2: t2.
Chow LM, Veillette A. 1995. The Src and Csk families of tyrosine protein kinases in hemopoietic cells.
Semin Immunol 7: 207–226.
Chu DH, Morita CT, Weiss A. 1998. The Syk family of protein tyrosine kinases in T-cell activation
and development. Immunol Rev 165: 167–180.
Daëron M, Lesourne R. 2006. Negative signaling in Fc receptor complexes. Adv Immunol 89: 39–86.
Das J, Ho M, Zikherman J, Govern C, Yang M, Weiss A, Chakraborty AK, Roose JP. 2009. Digital
signaling and hysteresis characterize ras activation in lymphoid cells. Cell 136: 337–351.
Davis MM. 2004. The evolutionary and structural “logic” of antigen receptor diversity. Semin Immunol
16: 239–243.
Depoil D, Fleire S, Treanor BL, Weber M, Harwood NE, Marchbank KL, Tybulewicz VL, Batista FD.
2008. CD19 is essential for B cell activation by promoting B cell receptor-antigen microcluster
formation in response to membrane-bound ligand. Nat Immunol 9: 63–72.
Dustin ML, Chakraborty AK, Shaw AS. 2010. Understanding the structure and function of the
immunological synapse. Cold Spring Harb Perspect Biol 2: a002311.
Elgueta R, Benson MJ, de Vries VC, Wasiuk A, Guo Y, Noelle RJ. 2009. Molecular mechanism and
function of CD40/CD40L engagement in the immune system. Immunol Rev 229: 152–172.
Finlay D, Cantrell D. 2011a. The coordination of T-cell function by serine/threonine kinases. Cold
Spring Harb Perspect Biol 3: a002261.
Finlay D, Cantrell D. 2011b. Metabolism, migration and memory in cytotoxic T cells. Nat Rev Immunol
11: 109–117.
Francisco LM, Sage PT, Sharpe AH. 2010. The PD-1 pathway in tolerance and autoimmunity. Immunol
Rev 236: 219–242.
Gallo EM, Canté-Barrett K, Crabtree GR. 2006. Lymphocyte calcium signaling from membrane to
nucleus. Nat Immunol 7: 25–32.
Gerondakis S, Siebenlist U. 2012. Roles of the NF-κB pathway in lymphocyte development and
function. Cold Spring Harb Perspect Biol 2: a000182.
Ghoreschi K, Laurence A, O’Shea JJ. 2009. Janus kinases in immune cell signaling. Immunol Rev 228:
273–287.
Harwood NE, Batista FD. 2011. The cytoskeleton coordinates the early events of B-cell activation.
Cold Spring Harb Perspect Biol 3: a002360.
* Harrison DA. 2012. The JAK/STAT pathway. Cold Spring Harb Perspect Biol 4: a011205.
Hayday AC. 2009. γδ T cells and the lymphoid stress-surveillance response. Immunity 31: 184–196.
Hedrick SM. 2009. The cunning little vixen: Foxo and the cycle of life and death. Nat Immunol 10:
1057–1063.
* Hemmings BA, Restuccia DF. 2012. PI3K-PKB/Akt pathway. Cold Spring Harb Perspect Biol 4:
a011189.
Hermiston ML, Zikherman J, Zhu JW. 2009. CD45, CD148, and Lyp/Pep: Critical phosphatases
regulating Src family kinase signaling networks in immune cells. Immunol Rev 228: 288–311.
Hogan PG, Lewis RS, Rao A. 2010. Molecular basis of calcium signaling in lymphocytes: STIM and
ORAI. Annu Rev Immunol 28: 491–533.
Im SH, Rao A. 2004. Activation and deactivation of gene expression by Ca2+/calcineurin-NFAT-
mediated signaling. Mol Cells 18: 1–9.
Jenkins MR, Griffiths GM. 2010. The synapse and cytolytic machinery of cytotoxic T cells. Curr Opin
Immunol 22: 308–313.
Jordan MS, Koretzky GA. 2010. Coordination of receptor signaling in multiple hematopoietic cell
lineages by the adaptor protein SLP-76. Cold Spring Harb Perspect Biol 2: a002501.
Kambayashi T, Larosa DF, Silverman MA, Koretzky GA. 2009. Cooperation of adapter molecules in
proximal signaling cascades during allergic inflammation. Immunol Rev 232: 99–114.
Karin M, Gallagher E. 2009. TNFR signaling: Ubiquitin-conjugated TRAFfic signals control stop-and-
go for MAPK signaling complexes. Immunol Rev 228: 225–240.
Kim ST, Takeuchi K, Sun ZY, Touma M, Castro CE, Fahmy A, Lang MJ, Wagner G, Reinherz EL.
2009. The αβ T cell receptor is an anisotropic mechanosensor. J Biol Chem 284: 31028–31037.
Kinashi T. 2005. Intracellular signalling controlling integrin activation in lymphocytes. Nat Rev
Immunol 5: 546–559.
Koretzky GA, Abtahian F, Silverman MA. 2006. SLP76 and SLP65: Complex regulation of signalling
in lymphocytes and beyond. Nat Rev Immunol 6: 67–78.
Krogsgaard M, Davis MM. 2005. How T cells “see” antigen. Nat Immunol 6: 239–245.
Kuo T, Schlissel MS. 2009. Mechanisms controlling expression of the RAG locus during lymphocyte
development. Curr Opin Immunol 21: 173–178.
Kurosaki T, Hikida M. 2009. Tyrosine kinases and their substrates in B lymphocytes. Immunol Rev
228: 132–148.
* Laplante M, Sabatini DM. 2012. mTOR signaling. Cold Spring Harb Perspect Biol 4: a011593.
* Lim K-H, Staudt LM. 2013. Toll-like receptor signaling. Cold Spring Harb Perspect Biol 5: a011247.
Lorenz U. 2009. SHP-1 and SHP-2 in T cells: Two phosphatases functioning at many levels. Immunol
Rev 228: 342–359.
Love PE, Hayes SM. 2010. ITAM-mediated signaling by the T-cell antigen receptor. Cold Spring Harb
Perspect Biol 2: a002485.
Macintyre AN, Finlay D, Preston G, Sinclair LV, Waugh CM, Tamas P, Feijoo C, Okkenhaug K,
Cantrell DA. 2011. Protein kinase B controls transcriptional programs that direct cytotoxic T cell
fate but is dispensable for T cell metabolism. Immunity 34: 224–236.
Mackay F, Figgett WA, Saulep D, Lepage M, Hibbs ML. 2010. B-cell stage and context-dependent
requirements for survival signals from BAFF and the B-cell receptor. Immunol Rev 237: 205–225.
Matthews SA, Cantrell DA. 2009. New insights into the regulation and function of serine/threonine
kinases in T lymphocytes. Immunol Rev 228: 241–252.
Matthews SA, Navarro MN, Sinclair LV, Emslie E, Feijoo-Carnero C, Cantrell DA. 2010. Unique
functions for protein kinase D1 and protein kinase D2 in mammalian cells. Biochem J 432: 153–
163.
Mocsai A, Ruland J, Tybulewicz VL. 2010. The SYK tyrosine kinase: A crucial player in diverse
biological functions. Nat Rev Immunol 10: 387–402.
* Morrison DK. 2012. MAP kinase pathways. Cold Spring Harb Perspect Biol 4: a011254.
Muller MR, Rao A. 2010. NFAT, immunity and cancer: A transcription factor comes of age. Nat Rev
Immunol 10: 645–656.
* Newton K, Dixit VM. 2012. Signaling in innate immunity and inflammation. Cold Spring Harb
Perspect Biol 4: a006049.
Nika K, Soldani C, Salek M, Paster W, Gray A, Etzensperger R, Fugger L, Polzella P, Cerundolo V,
Dushek O, et al. 2010. Constitutively active Lck kinase in T cells drives antigen receptor signal
transduction. Immunity 32: 766–777.
Nitschke L. 2009. CD22 and Siglec-G: B-cell inhibitory receptors with distinct functions. Immunol Rev
230: 128–143.
Oeckinghaus A, MS Hayden, Ghosh S. 2011. Crosstalk in NF-κB signaling pathways. Nat Immunol 12:
695–708.
Oellerich T, Bremes V, Neumann K, Bohnenberger H, Dittmann K, Hsiao HH, Engelke M, Schnyder
T, Batista FD, Urlaub H, et al. 2011. The B-cell antigen receptor signals through a preformed
transducer module of SLP65 and CIN85. EMBO J 30: 3620–3634.
Oh-hora M, Rao A. 2008. Calcium signaling in lymphocytes. Curr Opin Immunol 20: 250–258.
Okkenhaug K, Fruman DA. 2010. PI3Ks in lymphocyte signaling and development. Curr Top
Microbiol Immunol 346: 57–85.
O’Shea JJ, Paul WE. 2010. Mechanisms underlying lineage commitment and plasticity of helper CD4+
T cells. Science 327: 1098–1102.
Parry RV, Harris SJ, Ward SG. 2010. Fine tuning T lymphocytes: A role for the lipid phosphatase
SHIP-1. Biochim Biophys Acta 1804: 592–597.
Powell JD, Delgoffe GM. 2010. The mammalian target of rapamycin: Linking T cell differentiation,
function, and metabolism. Immunity 33: 301–311.
Quann EJ, Merino E, Furuta T, Huse M. 2009. Localized diacylglycerol drives the polarization of the
microtubule-organizing center in T cells. Nat Immunol 6: 627–635.
Qureshi OS, Zheng Y, Nakamura K, Attridge K, Manzotti C, Schmidt EM, Baker J, Jeffery LE, Kaur S,
Briggs Z, et al. 2011. Trans-endocytosis of CD80 and CD86: A molecular basis for the cell-
extrinsic function of CTLA-4. Science 332: 600–603.
Rochman Y, Spolski R, Leonard WJ. 2009. New insights into the regulation of T cells by γc family
cytokines. Nat Rev Immunol 9: 480–490.
Rudd CE. 2008. The reverse stop-signal model for CTLA4 function. Nat Rev Immunol 8: 153–160.
Salmond RJ, Filby A, Qureshi I, Caserta S, Zamoyska R. 2009. T-cell receptor proximal signaling via
the Src-family kinases, Lck and Fyn, influences T-cell activation, differentiation, and tolerance.
Immunol Rev 228: 9–22.
* Samelson LE. 2011. Immunoreceptor signaling. Cold Spring Harb Perspect Biol 3: a011510.
Schatz DG, Ji Y. 2011. Recombination centres and the orchestration of V(D)J recombination. Nat Rev
Immunol 11: 251–263.
Sinclair LV, Finlay D, Feijoo C, Cornish GH, Gray A, Ager A, Okkenhaug K, Hagenbeek TJ, Spits H,
Cantrell DA. 2008. Phosphatidylinositol-3-OH kinase and nutrient-sensing mTOR pathways control
T lymphocyte trafficking. Nat Immunol 9: 513–521.
Spitaler M, Emslie E, Wood CD, Cantrell D. 2006. Diacylglycerol and protein kinase D localization
during T lymphocyte activation. Immunity 24: 535–546.
Springer TA, Dustin ML. 2011. Integrin inside-out signaling and the immunological synapse. Curr
Opin Cell Biol 24: 107–115.
Srinivasan L, Sasaki Y, Calado DP, Zhang B, Paik JH, DePinho RA, Kutok JL, Kearney JF, Otipoby
KL, Rajewsky K. 2009. PI3 kinase signals BCR-dependent mature B cell survival. Cell 139: 573–
586.
Tolar P, Sohn HW, Liu W, Pierce SK. 2009. The molecular assembly and organization of signaling
active B-cell receptor oligomers. Immunol Rev 232: 34–41.
van der Merwe PA, Davis SJ. 2003. Molecular interactions mediating T cell antigen recognition. Annu
Rev Immunol 21: 659–684.
van der Merwe PA, Dushek O. 2011. Mechanisms for T cell receptor triggering. Nat Rev Immunol 11:
47–55.
Veillette A. 2010. SLAM-family receptors: Immune regulators with or without SAP-family adaptors.
Cold Spring Harb Perspect Biol 2: a002469.
Veillette A, Latour S, Davidson D. 2002. Negative regulation of immunoreceptor signaling. Annu Rev
Immunol 20: 669–707.
Walker LS, Sansom DM. 2011. The emerging role of CTLA4 as a cell-extrinsic regulator of T cell
responses. Nat Rev Immunol 11: 852–863.
Wang H, Kadlecek TA, Au-Yeung BB, Goodfellow HE, Hsu LY, Freedman TS, Weiss A. 2010. ZAP-
70: An essential kinase in T-cell signaling. Cold Spring Harb Perspect Biol 2: a002279.
* Ward PS, Thompson CB. 2012. Signaling in control of cell growth and metabolism. Cold Spring
Harb Perspect Biol 4: a006783.
Watts TH. 2005. TNF/TNFR family members in costimulation of T cell responses. Annu Rev Immunol
23: 23–68.
Xiong N, Raulet DH. 2007. Development and selection of γδ T cells. Immunol Rev 215: 15–31.
Yang J, Reth M. 2010a. The dissociation activation model of B cell antigen receptor triggering. FEBS
Lett 584: 4872–4877.
Yang J, Reth M. 2010b. Oligomeric organization of the B-cell antigen receptor on resting cells. Nature
467: 465–469.
Zikherman J, Jenne C, Watson S, Doan K, Raschke W, Goodnow CC, Weiss A. 2010. CD45-Csk
phosphatase-kinase titration uncouples basal and inducible T cell receptor signaling during thymic
development. Immunity 32: 342–354.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a018788
CHAPTER 17
Vertebrate Reproduction
SUMMARY
Outline
1 Introduction
2 Oocyte maturation
3 Sperm maturation
4 Fertilization
5 From egg to zygote
6 Concluding remarks
References
1 INTRODUCTION
Mammalian reproduction depends on the proper development and maturation
of both the female egg and the male sperm. These gametes fuse through a
complex series of events, known as fertilization, that ensure the highest
quality of offspring. Both gamete development and fertilization depend on
numerous connected signaling pathways, a flaw in any of which can lead to
infertility or birth defects.
The egg and sperm are haploid germ cells that, upon fertilization,
reconstitute a diploid cell—the embryo. Production of haploid gametes from
diploid precursors requires a modified cell cycle known as meiosis. Before
meiosis, the full complement of parental chromosomes is first duplicated in S
phase, to produce so-called sister chromatids (i.e., four copies of each
chromosome per cell) and paternal and maternal chromosomes pair up. The
homologous chromosomes from each chromosome pair are then separated in
the first meiotic M phase (meiosis I, also known as MI). Subsequently,
without further replication, the cells reenter M phase (meiosis II, also known
as MII) to divide the sister chromatids equally into four haploid daughter
cells. For male gametes, four mature sperm are generated, whereas in the
female, a single final gamete (the egg) is produced together with three polar
bodies. Movement through these stages of meiosis is carefully controlled by
kinases, phosphatases, ubiquitin-dependent degradation of key regulators,
and calcium flux.
Before acquiring the capacity to fertilize eggs, a sperm must reside in the
female reproductive tract and undergo physiological changes that render it
fertilization competent (Bedford 1970). The acquisition of fertilization
competence and the biochemical, membrane, and enzymatic changes that
underlie it are collectively known as capacitation (Austin 1951; Chang 1951).
As with gamete development and maturation, capacitation and fertilization
depend on careful regulation through signaling pathways. These include
pathways involving gonadotropins, G-protein-coupled receptors (GPCRs),
kinases, and calcium signaling (Salicioni et al. 2007).
2 OOCYTE MATURATION
Oocyte maturation has been most extensively studied in the frog Xenopus
laevis, because its very large oocytes allow both physical manipulation of the
cell (microinjection of proteins, RNAs, and antisense oligonucleotides) and
observation of the progression through meiosis with the naked eye. Although
some notable differences have been observed, genetic studies in mammals
(primarily mouse) have revealed similar overall regulation of meiotic
progression (Fig. 1).
Figure 1. Regulation of oocyte maturation. Immature oocytes are held in G2 arrest through the activity
of PKA, which is stimulated by GPR3-dependent production of cAMP. Progesterone signaling leads to
a loss of PKA activity, leading to disinhibition of the CDK1 activator Cdc25 and stimulation of the
CDK1 inhibitor Wee1. Additionally, progesterone stimulates translation of both Mos and cyclin B
proteins (the former also stimulating translation of the latter) through induction of Eg2, which
phosphorylates and activates CPEB to unmask the messages and promote their polyadenylation. In
parallel, progesterone stimulates translation of RINGO, which can activate CDK1 independently of
cyclin and make it phosphorylate and inhibit the CDK1 inhibitor Myt1. All of these events result in
activation of MPF, which drives the oocyte into MI. Maturation continues via the Mos-MEK-ERK
pathway as shown. Blockade of APC (anaphase-promoting complex) activity by CSF and Emi2 holds
the mature oocyte at a second arrest until the time of fertilization.
Oocytes and sperm both begin life as primordial germ cells (PGCs) that
migrate to the nascent gonads (ovaries in females, testes in males) in early
embryonic development. Under the influence of a variety of cytokines and
growth factors, PGCs that will become oocytes continue dividing mitotically
within cell clusters. In oogenesis, the premeiotic S phase is followed by a
prolonged arrest in prophase I of meiosis until sexual maturity. During this
phase, the oocyte is maintained in a G2-phase-arrested state through G-
protein-coupled signaling (see below). When mitosis ceases, these oocytes
each become surrounded by somatic granulosa and theca cells, which form
the primordial follicles that serve as repositories of dormant oocytes for later
ovulation. Oocytes nestled within the follicles grow and stockpile nutrients
until they become competent to undergo maturation; upon receipt of
appropriate hormonal signals, one follicle from the larger pool will mature
fully during each menstrual cycle in the mammal. Stimulated by pituitary
hormones (gonadotropins) and as a consequence of maturation-inducing
steroid hormones (e.g., progesterone) synthesized by the ovarian follicle
cells, the oocyte exits prophase arrest, and progresses through MI,
transitioning promptly to MII without any intervening DNA replication. At
MII, the oocyte arrests again awaiting fertilization.
The end product of oocyte maturation is a haploid egg capable of being
fertilized. A strong MII arrest helps to prevent parthenogenesis, which is the
aberrant entry of the haploid egg into the mitotic cell cycle in the absence of
fertilization. In an effort to define the factors responsible for MII arrest,
Masui and colleagues injected extract prepared from a mature M-phase-
arrested frog egg into blastomeres formed after the first embryonic cell
division (Masui and Markert 1971); injected cells remained arrested in M
phase, whereas the uninjected cells continued to divide. These experiments
helped to identify both maturation promoting factor (MPF), which drives
entry into both MI and MII during oocyte development, and the cytostatic
factor (CSF), which maintains MII arrest. We now know that MPF is
equivalent to the complex of cyclin B and cyclin-dependent kinase 1 (cyclin-
B–CDK1) that drives entry into mitosis in the somatic cell cycle (Dunphy et
al. 1988; Labbe et al. 1989; Ch. 6 [Rhind and Russell 2012]). CSF was shown
to be the kinase Mos, the cellular counterpart of the viral oncoprotein v-Mos,
which is expressed primarily in germ cells (Propst et al. 1987; Sagata et al.
1989). Proper maturation from MI entry through MII arrest depends on
tightly controlled temporal regulation of both cyclin-B–CDK1 and Mos
activity.
2.1 Meiosis I
During MI, oocytes are maintained in the G2-arrested state by high levels of
cytosolic cAMP. Constitutive signaling by the GPCR GPR3 probably
stimulates adenylyl cyclase to keep cAMP levels high. Indeed,
overexpression of GPR3 in frog oocytes makes them resistant to the
maturation effects of progesterone, and in mice lacking GPR3, oocytes
mature in the absence of additional stimuli (Freudzon et al. 2005; Hinckley et
al. 2005; Mehlmann 2005; Deng et al. 2008). Although sphingosine 1-
phosphate and sphingosylphosphorylcholine have been proposed as GPR3
ligands, unliganded GPR3 appears to be able to stimulate adenylyl cyclase
and thus the role of GPR3 in maintenance of G2 arrest is not entirely clear
(Eggerickx et al. 1995; Uhlenbrock et al. 2002; Hinckley et al. 2005).
Progesterone stimulation seems to antagonize the GPR3 signal, triggering a
decrease in cAMP levels that is at least partly mediated by stimulation of
phosphodiesterases that degrade cAMP (primarily PDE3 in the oocytes)
(Tsafriri et al. 1996). This leads to a diminution of protein kinase A (PKA)
activity. If PDE3 is artificially inhibited or cAMP synthesis is artificially
stimulated, progesterone-induced maturation can be blocked. Conversely,
injection of oocytes with the PKA inhibitor PKI can promote resumption of
meiosis even without progesterone stimulation (Stanford et al. 2003).
As in the mitotic cell cycle, cyclin-B–CDK1 activity is controlled by
phosphorylation of CDK1 on Y15 by Wee1-family kinases, which is opposed
by the Cdc25 phosphatase (Watanabe et al. 1995; Berry and Gould 1996).
The Wee1 relative Myt1 contributes to suppressing CDK1 phosphorylation,
and the Cdc25c isoform mediates its subsequent dephosphorylation. In mice,
this process is mediated by oocyte-specific isoforms: WEE1B and CDC25B.
Loss of WEE1B irreversibly arrests oocytes in prophase (Han et al. 2005). In
the G2-arrested oocyte, PKA directly phosphorylates Cdc25 on S287
(Xenopus numbering), promoting the binding of the small acidic protein 14-
3-3 (Duckworth et al. 2002). 14-3-3 interferes with the ability of Cdc25 to
interact with and dephosphorylate cyclin-B–CDK1 and prevents its
translocation into the nucleus, where it would promote rapid cyclin-B–CDK1
activation (Kumagai and Dunphy 1999; Lopez-Girona et al. 1999; Yang et al.
1999). Thus, the drop in PKA activity required for maturation promotes
Cdc25 activation. PKA also phosphorylates and activates Wee1/Myt1
(Stanford and Ruderman 2005); so the drop in PKA also promotes CDK1
activation by alleviating its suppression by these kinases.
The formation and activity of MPF is also regulated by translation of
cyclin B, whose mRNA is translationally dormant before the induction of
oocyte maturation owing to its very short poly-A tail. At the time of oocyte
maturation, cis-acting sequences within the 3′ UTR of the cyclin B mRNA
promote cytoplasmic polyadenylation, elongating the tail more than 100
nucleotides. These cytoplasmic polyadenylation elements (CPEs) within the
3′ UTR of the mRNA are bound by CPE-binding protein (CPEB) (Hake and
Richter, 1994). Through a process that is not entirely clear, the drop in PKA
activity that heralds the onset of oocyte maturation also induces activation of
a kinase, Eg2, which phosphorylates CPEB, activating it to both unmask the
mRNA and recruit a poly(A) polymerase to elongate the poly(A) tail
(Andresson and Ruderman 1998; Frank-Vaillant et al. 2000; Hodgman et al.
2001).
Note that in every species there is at least some preformed cyclin-B–
CDK1 complex (known as pre-MPF) whose activity is suppressed by
phosphorylation of CDK1 at T14 and Y15. Indeed, the initial discovery of
MPF relied on the ability of the injected MPF to mobilize the pre-MPF pool
through autoamplification (Masui and Markert 1971; Drury and Schorderet-
Slatkine 1975; Wasserman and Masui 1975). Phosphorylation by cyclin-B–
CDK1 suppresses Wee1/Myt1 and activates Cdc25, which promotes more
conversion of pre-MPF to MPF. In species in which most of the CDK1 is
bound to cyclin B in pre-MPF complexes, new cyclin B translation is not
absolutely required for induction of oocyte maturation; however, in those
species that have low amounts of pre-MPF and high levels of free CDK1,
cyclin B synthesis is an obligate step in maturation (Jagiello 1969; Fulka et
al. 1986; Moor and Crosby 1986; Hunter and Moor 1987; Gautier and Maller
1991; Mattioli et al. 1991).
In mammals, the signal emanating from a loss of PKA activity may be
conveyed directly to CDK1 via Wee1B, as this appears to be a direct PKA
target. For Myt1, there is evidence for indirect pathways of inhibition. First,
soon after progesterone treatment, a non-cyclin alternative activator of CDK1
known as RINGO is translated (Ferby et al. 1999). This protein can bind to
and activate CDK1, causing it to phosphorylate and suppress Myt1 (Ruiz et
al. 2008). This, in turn, leads to activation of cyclin-B–CDK1 complexes. A
second pathway is an oocyte-specific MAP kinase (MAPK) cascade
involving the MAPKKK Mos, the MAPKK MEK, and the MAPK ERK. In
frogs, the terminal effector in this pathway is RSK, which can phosphorylate
and inhibit Myt1 (Palmer et al. 1998). RSK can also phosphorylate Cdc25,
contributing to its activation. Accordingly, injection of activated RSK into
Xenopus oocytes can induce meiotic maturation and RSK inhibition interferes
with progesterone-induced maturation. In mice, alternative pathways must
operate (e.g., the direct inhibition of WEE1B by PKA, as described above),
because mice lacking all known isoforms of RSK do not show defects in
oocyte maturation (Dumont et al. 2005).
In some species, including Xenopus, activation of the Mos-ERK pathway
precedes completion of MI and breakdown of the nuclear envelope (known as
germinal vesicle breakdown [GVBD] in oocytes). In others, it occurs after
GVBD (because cyclin-B–CDK1 can actually activate ERK). Whether the
Mos-MEK-ERK-RSK pathway is involved at MI depends on the organism,
and Mos accumulation is controlled by multiple mechanisms (reviewed in
Fan and Sun 2004). The 3′ end of the Mos mRNA in the immature oocyte has
a short poly(A) tail whose elongation (necessary for efficient translation) is
masked through binding of CPEB. In addition to the Eg2-induced
phosphorylation of CPEB, which unmasks and enhances the translation of
Mos (Mendez et al. 2000), the stability of Mos protein is greatly enhanced by
phosphorylation at S3 as both dephosphorylation of this residue and the
presence of a proline at residue 2 are required for recognition by the
ubiquitin-proteasome degradation system (Nishizawa et al. 1993). This site is
phosphorylated in a positive-feedback loop by ERK, which stabilizes Mos.
Where Mos accumulates only after GVBD, it is phosphorylated at the same
site by cyclin-B–CDK1. Together, increased translation and stabilization
promote Mos accumulation during maturation. Note that, in Xenopus,
redundant pathways allow MI progression even when Mos is ablated; yet, the
kinetics are delayed, indicating that Mos normally enhances meiotic
progression in this species. Indeed, when either cyclin B synthesis or Mos
synthesis is impaired, progesterone-induced GVBD can proceed, but ablation
of both abolishes this.
Figure 2. Regulation of Emi2 and the APC during the MI–MII transition. Phosphorylation controls
Emi2 stability during oocyte maturation. At MI, CDK1 phosphorylates four amino-terminal sites
(S213, T239, T252, and T267) on Emi2; this triggers Emi2 degradation, required for MI exit. At MI
anaphase, cyclin B is degraded, leading to a drop in CDK1 activity. Emi2 is stabilized by
dephosphorylation triggered by Mos signaling. Emi accumulates, resulting in APC inhibition, critical
for S phase block and MII entry. CDK1 activity is low in MII relative to MI. Emi2 is stable in MII, as
required for CSF arrest. At fertilization, Emi2 is quickly degraded through a CaMKII-mediated
pathway, allowing activation of the APC and exit from MII. At the onset of MI anaphase, APC-
mediated cyclin B degradation results in decreased CDK1 activity. With the Mos-PP2A pathway
predominant, dephosphorylated and stabilized Emi2 protein prevents complete ubiquitylation of cyclin
B by the APC. This is essential for the inhibition of S phase between MI and MII. (From Tang et al.
2008; adapted, with permission.)
3 SPERM MATURATION
Spermatogenesis occurs over the course of several weeks and encompasses
three successive phases (Sharpe 1994): proliferation, meiosis, and
differentiation. During proliferation, spermatogonial stem cells (SSCs)
differentiate into spermatogonia. These undergo several mitotic divisions,
giving rise to spermatocytes. After two meiotic divisions, spermatocytes form
haploid spermatids. The final transformation of spermatids into mature sperm
entails a major physical and structural reorganization of the cell that is known
as spermiogenesis (Fig. 3).
Figure 3. Regulation of sperm maturation. (A) Spermatogenesis is a continuous process that starts after
puberty. This is possible because a subset of spermatogonium, spermatogonial stem cells (SSCs), are
capable of self-renewal. Glial-cell-line-derived neurotrophic factor (GDNF) and the receptor complex
composed of Ret protooncoprotein and the GDNF family receptor α1 (GFRα1) are key signaling events
for self-renewal. (From Hofmann 2008; adapted, with permission.) (B) Spermatocytes undergo a series
of maturation steps before differentiating into sperm. These are tightly regulated by interaction with
Sertoli cells and testis factors such as GDNF, ERM, KitL, and retinoic acid (RA). (C) Sperm must
additionally undergo capacitation before they are capable of fertilization. This includes both fast and
slow events. The fast events stimulate flagellar activity through a calcium flux that is controlled by the
adenylyl cyclase SACY and mediated by the CatSper and sodium/bicarbonate (NBC) channels. Slow
events increase motility and introduce changes that prepare sperm for fertilization. These include
increases in both intracellular calcium levels and tyrosine phosphorylation of PKA substrates. cAMP
levels are also controlled by endogenous phosphodiesterase (PDE). Unknown cholesterol acceptors
present in uterine/oviduct fluids mediate cholesterol efflux during capacitation. (From Visconti 2009;
adapted, with permission.)
4 FERTILIZATION
4.1 The Acrosome Reaction in Sperm
Before fusing with the egg’s plasma membrane, sperm must undergo the
acrosome reaction (Yanagimachi and Mahi 1976). The acrosome is a Golgi-
derived organelle that lies above the tip of the sperm head. Its contents are
released following fusion between the outer acrosomal membrane and the
plasma membrane (Kim et al. 2011), and this reaction is critical for
interaction with the ZP of the egg. Only capacitated sperm are capable of
undergoing the acrosome reaction. Progesterone produced by the cumulus
cells surrounding the oocyte has been proposed as the possible inducer of this
reaction (Osman et al. 1989; Jin et al. 2011). Upon breaching the egg’s
plasma membrane, the sperm induces the initiation of embryonic
development by evoking an increase in the intracellular concentration of free
calcium, a signaling mechanism that regulates numerous cellular processes
(Fig. 4) (Berridge et al. 2000b; p. 95 [Bootman 2012]).
Figure 4. Fertilization events. Both egg and sperm must undergo complex changes before fertilization
can occur. Upon interaction with the zona pelucida, sperm undergo a calcium-flux-driven
reorganization of SNARE fusion proteins, called the acrosome reaction. Fusion with the egg membrane
releases factors from the sperm into the cytoplasm. These factors stimulate calcium release from the
egg ER and subsequent activation of CaMKII and calcineurin (CN). The consequences of CaMKII
activation include phosphorylation of Emi2, which promotes its Plx1 and SCFβ-TrCP-dependent
degradation and consequent reactivation of the APC. CN activation promotes dephosphorylation of
Cdc20 and also the Cdc27 subunit of the APC. At the same time, dephosphorylation of M phase
phosphoproteins is promoted by inactivation of the Greatwall kinase (Gwl), thereby alleviating
inhibition of PP2A, allowing it to dephosphorylate M phase CDK1 substrates.
6 CONCLUDING REMARKS
In gametes of both sexes, development, maturation, and fertilization require
the careful coordination of complex processes. From hormone-initiated and
kinase/phosphatase-controlled maturation, to calcium-induced capacitation
and fertilization, regulatory mechanisms ensure that reproduction occurs only
under conditions in which they are best poised for success. For males, the
signals that regulate spermatogenesis are at first contained within the testis,
but following spermiation and ejaculation, proper sperm function depends on
factors outside of the male reproductive tract: for external fertilizers, factors
in the outside environment; and for internal fertilizers, the milieu of the
female reproductive tract. Taking into account these potentially harsh
environments, male reproduction relies on the production of large quantities
of sperm, with progenitor cells retaining mitotic capacity into adulthood. For
female internal fertilizers, the production of eggs is more contained, restricted
to the follicle until ovulation and to the female ducts until fertilization and
beyond, relying on the production of few gametes; in contrast, externally
fertilizing females, like their male counterparts, produce large numbers of
gametes. Even for these organisms, though, oocytes and eggs are self-
contained developmental units capable of sustaining the early stages of
development whose progenitors have generally entered meiosis and have lost
the capacity to regenerate.
In every species, regardless of the site of fertilization, cascades of protein
modifications regulate cell-cycle transitions to guarantee that oocytes can be
fertilized only at an appropriate time. Hormonal regulation and calcium
signaling promote capacitation of sperm only in the correct environment for
fertilization. Furthermore, multiple overlapping pathways provide the checks
and balances that are necessary to prevent defective reproduction.
Although great progress has been made over the last 50 years in detailing
the molecular events underlying most aspects of vertebrate fertilization, there
are still aspects of gamete development and fertilization whose precise
regulation by cell signaling events remain to be determined. Elucidation of
these events is likely to have important implications for the continued
development of reproductive technologies and for maximizing the health of
gametes, and thus of progeny.
REFERENCES
*Reference is in this book.
Andresson T, Ruderman JV. 1998. The kinase Eg2 is a component of the Xenopus oocyte progesterone-
activated signaling pathway. EMBO J 17: 5627–5637.
Araki K, Naito K, Haraguchi S, Suzuki R, Yokoyama M, Inoue M, Aizawa S, Toyoda Y, Sato E. 1996.
Meiotic abnormalities of c-mos knockout mouse oocytes: Activation after first meiosis or entrance
into third meiotic metaphase. Biol Reprod 55: 1315–1324.
Arnoult C, Cardullo RA, Lemos JR, Florman HM. 1996. Activation of mouse sperm T-type Ca2+
channels by adhesion to the egg zona pellucida. Proc Natl Acad Sci 93: 13004–13009.
Austin CR. 1951. Observations on the penetration of the sperm in the mammalian egg. Aust J Sci Res B
4: 581–596.
Avella MA, Xiong B, Dean J. 2013. The molecular basis of gamete recognition in mice and humans.
Mol Hum Reprod 19: 279–289.
Baba T, Azuma S, Kashiwabara S, Toyoda Y. 1994. Sperm from mice carrying a targeted mutation of
the acrosin gene can penetrate the oocyte zona pellucida and effect fertilization. J Biol Chem 269:
31845–31849.
Bedell MA, Mahakali Zama A. 2004. Genetic analysis of Kit ligand functions during mouse
spermatogenesis. J Androl 25: 188–199.
Bedford JM. 1970. Sperm capacitation and fertilization in mammals. Biol Reprod 2: 128–158.
Bedford-Guaus SJ, Yoon SY, Fissore RA, Choi YH, Hinrichs K. 2008. Microinjection of mouse
phospholipase Cζ complementary RNA into mare oocytes induces long-lasting intracellular calcium
oscillations and embryonic development. Reprod Fertil Dev 20: 875–883.
Berridge MJ, Lipp P, Bootman MD. 2000a. Signal transduction. The calcium entry pas de deux.
Science 287: 1604–1605.
Berridge MJ, Lipp P, Bootman MD. 2000b. The versatility and universality of calcium signalling. Nat
Rev Mol Cell Biol 1: 11–21.
Berry LD, Gould KL. 1996. Regulation of Cdc2 activity by phosphorylation at T14/Y15. Prog Cell
Cycle Res 2: 99–105.
Bohmer M, Van Q, Weyand I, Hagen V, Beyermann M, Matsumoto M, Hoshi M, Hildebrand E, Kaupp
UB. 2005. Ca2+ spikes in the flagellum control chemotactic behavior of sperm. EMBO J 24: 2741–
2752.
* Bootman MD. 2012. Calcium signaling. Cold Spring Harb Perspect Biol 4: a011171.
Bootman MD, Collins TJ, Peppiatt CM, Prothero LS, MacKenzie L, De Smet P, Travers M, Tovey SC,
Seo JT, Berridge MJ, et al. 2001. Calcium signalling—An overview. Sem Cell Dev Biol 12: 3–10.
Bowles J, Knight D, Smith C, Wilhelm D, Richman J, Mamiya S, Yashiro K, Chawengsaksophak K,
Wilson MJ, Rossant J, et al. 2006. Retinoid signaling determines germ cell fate in mice. Science
312: 596–600.
Buaas FW, Kirsh AL, Sharma M, McLean DJ, Morris JL, Griswold MD, de Rooij DG, Braun RE.
2004. Plzf is required in adult male germ cells for stem cell self-renewal. Nat Genet 36: 647–652.
Carlson AE, Westenbroek RE, Quill T, Ren D, Clapham DE, Hille B, Garbers DL, Babcock DF. 2003.
CatSper1 required for evoked Ca2+ entry and control of flagellar function in sperm. Proc Natl Acad
Sci 100: 14864–14868.
Chang MC. 1951. Fertilizing capacity of spermatozoa deposited into the fallopian tubes. Nature 168:
697–698.
Chang H, Suarez SS. 2010. Rethinking the relationship between hyperactivation and chemotaxis in
mammalian sperm. Biol Reprod 83: 507–513.
Chang C, Chen YT, Yeh SD, Xu Q, Wang RS, Guillou F, Lardy H, Yeh S. 2004. Infertility with
defective spermatogenesis and hypotestosteronemia in male mice lacking the androgen receptor in
Sertoli cells. Proc Natl Acad Sci 101: 6876–6881.
Chen Y, Cann MJ, Litvin TN, Iourgenko V, Sinclair ML, Levin LR, Buck J. 2000. Soluble adenylyl
cyclase as an evolutionarily conserved bicarbonate sensor. Science 289: 625–628.
Chen C, Ouyang W, Grigura V, Zhou Q, Carnes K, Lim H, Zhao GQ, Arber S, Kurpios N, Murphy TL,
et al. 2005. ERM is required for transcriptional control of the spermatogonial stem cell niche.
Nature 436: 1030–1034.
Cheng J, Watkins SC, Walker WH. 2007. Testosterone activates mitogen-activated protein kinase via
Src kinase and the epidermal growth factor receptor in sertoli cells. Endocrinology 148: 2066–
2074.
Colledge WH, Carlton MB, Udy GB, Evans MJ. 1994. Disruption of c-mos causes parthenogenetic
development of unfertilized mouse eggs. Nature 370: 65–68.
Costoya JA, Hobbs RM, Barna M, Cattoretti G, Manova K, Sukhwani M, Orwig KE, Wolgemuth DJ,
Pandolfi PP. 2004. Essential role of Plzf in maintenance of spermatogonial stem cells. Nat Genet
36: 653–659.
Daar I, Paules RS, Vande Woude GF. 1991. A characterization of cytostatic factor activity from
Xenopus eggs and c-mos-transformed cells. J Cell Biol 114: 329–335.
De Gendt K, Swinnen JV, Saunders PT, Schoonjans L, Dewerchin M, Devos A, Tan K, Atanassova N,
Claessens F, Lecureuil C, et al. 2004. A Sertoli cell-selective knockout of the androgen receptor
causes spermatogenic arrest in meiosis. Proc Natl Acad Sci 101: 1327–1332.
Deng J, Lang S, Wylie C, Hammes SR. 2008. The Xenopus laevis isoform of G protein-coupled
receptor 3 (GPR3) is a constitutively active cell surface receptor that participates in maintaining
meiotic arrest in X. laevis oocytes. Mol Endocrinol 22: 1853–1865.
de Rooij DG. 2009. The spermatogonial stem cell niche. Microsc Res Tech 72: 580–585.
Díaz-Muñoz M, de la Rosa Santander P, Juárez-Espinosa AB, Arellano RO, Morales-Tlalpan V. 2008.
Granulosa cells express three inositol 1,4,5-trisphosphate receptor isoforms: Cytoplasmic and
nuclear Ca2+ mobilization. Reprod Biol Endocrinol 6: 60.
Distelhorst CW, Bootman MD. 2011. Bcl-2 interaction with the inositol 1,4,5-trisphosphate receptor:
Role in Ca2+ signaling and disease. Cell Calcium 50: 234–241.
Drury KC, Schorderet-Slatkine S. 1975. Effects of cycloheximide on the “autocatalytic” nature of the
maturation promoting factor (MPF) in oocytes of Xenopus laevis. Cell 4: 269–274.
Ducibella T, Huneau D, Angelichio E, Xu Z, Schultz RM, Kopf GS, Fissore R, Madoux S, Ozil JP.
2002. Egg-to-embryo transition is driven by differential responses to Ca2+ oscillation number. Dev
Biol 250: 280–291.
Duckworth BC, Weaver JS, Ruderman JV. 2002. G2 arrest in Xenopus oocytes depends on
phosphorylation of cdc25 by protein kinase A. Proc Natl Acad Sci 99: 16794–16799.
Dumont J, Umbhauer M, Rassinier P, Hanauer A, Verlhac MH. 2005. p90RSK is not involved in
cytostatic factor arrest in mouse oocytes. J Cell Biol 169: 227–231.
Dunphy WG, Brizuela L, Beach D, Newport J. 1988. The Xenopus cdc2 protein is a component of
MPF, a cytoplasmic regulator of mitosis. Cell 54: 423–431.
Dupre A, Jessus C, Ozon R, Haccard O. 2002. Mos is not required for the initiation of meiotic
maturation in Xenopus oocytes. EMBO J 21: 4026–4036.
Eggerickx D, Denef JF, Labbe O, Hayashi Y, Refetoff S, Vassart G, Parmentier M, Libert F. 1995.
Molecular cloning of an orphan G-protein-coupled receptor that constitutively activates adenylate
cyclase. Biochem J 309: 837–843.
Evans JP, Florman HM. 2002. The state of the union: The cell biology of fertilization. Nat Cell Biol 4:
S57–S63.
Fan HY, Sun QY. 2004. Involvement of mitogen-activated protein kinase cascade during oocyte
maturation and fertilization in mammals. Biol Reprod 70: 535–547.
Feng LX, Ravindranath N, Dym M. 2000. Stem cell factor/c-kit up-regulates cyclin D3 and promotes
cell cycle progression via the phosphoinositide 3-kinase/p70 S6 kinase pathway in spermatogonia. J
Biol Chem 275: 25572–25576.
Ferby I, Blazquez M, Palmer A, Eritja R, Nebreda AR. 1999. A novel p34cdc2-binding and activating
protein that is necessary and sufficient to trigger G2/M progression in Xenopus oocytes. Genes Dev
13: 2177–2189.
Feske S, Gwack Y, Prakriya M, Srikanth S, Puppel SH, Tanasa B, Hogan PG, Lewis RS, Daly M, Rao
A. 2006. A mutation in Orai1 causes immune deficiency by abrogating CRAC channel function.
Nature 441: 179–185.
Ficarro S, Chertihin O, Westbrook VA, White F, Jayes F, Kalab P, Marto JA, Shabanowitz J, Herr JC,
Hunt DF, et al. 2003. Phosphoproteome analysis of capacitated human sperm. Evidence of tyrosine
phosphorylation of a kinase-anchoring protein 3 and valosin-containing protein/p97 during
capacitation. J Biol Chem 278: 11579–11589.
Finch EA, Turner TJ, Goldin SM. 1991. Subsecond kinetics of inositol 1,4,5-trisphosphate-induced
calcium release reveal rapid potentiation and subsequent inactivation by calcium. Ann NY Acad Sci
635: 400–403.
Fissore RA, Long CR, Duncan RP, Robl JM. 1999a. Initiation and organization of events during the
first cell cycle in mammals: Applications in cloning. Cloning 1: 89–100.
Fissore RA, Longo FJ, Anderson E, Parys JB, Ducibella T. 1999b. Differential distribution of inositol
trisphosphate receptor isoforms in mouse oocytes. Biol Reprod 60: 49–57.
Flanagan JG, Chan DC, Leder P. 1991. Transmembrane form of the kit ligand growth factor is
determined by alternative splicing and is missing in the Sld mutant. Cell 64: 1025–1035.
Frank-Vaillant M, Haccard O, Thibier C, Ozon R, Arlot-Bonnemains Y, Prigent C, Jessus C. 2000.
Progesterone regulates the accumulation and the activation of Eg2 kinase in Xenopus oocytes. J
Cell Sci 113: 1127–1138.
Freudzon L, Norris RP, Hand AR, Tanaka S, Saeki Y, Jones TL, Rasenick MM, Berlot CH, Mehlmann
LM, Jaffe LA. 2005. Regulation of meiotic prophase arrest in mouse oocytes by GPR3, a
constitutive activator of the Gs G protein. J Cell Biol 171: 255–265.
Fujimoto S, Yoshida N, Fukui T, Amanai M, Isobe T, Itagaki C, Izumi T, Perry AC. 2004. Mammalian
phospholipase Cζ induces oocyte activation from the sperm perinuclear matrix. Dev Biol 274: 370–
383.
Fukami K, Nakao K, Inoue T, Kataoka Y, Kurokawa M, Fissore RA, Nakamura K, Katsuki M,
Mikoshiba K, Yoshida N, et al. 2001. Requirement of phospholipase Cδ4 for the zona pellucida-
induced acrosome reaction. Science 292: 920–923.
Fulka J Jr, Motlik J, Fulka J, Crozet N. 1986. Activity of maturation promoting factor in mammalian
oocytes after its dilution by single and multiple fusions. Dev Biol 118: 176–181.
Gautier J, Maller JL. 1991. Cyclin B in Xenopus oocytes: Implications for the mechanism of pre-MPF
activation. EMBO J 10: 177–182.
Giusti AF, Carroll DJ, Abassi YA, Foltz KR. 1999. Evidence that a starfish egg Src family tyrosine
kinase associates with PLC-γ1 SH2 domains at fertilization. Dev Biol 208: 189–199.
Goud PT, Goud AP, Leybaert L, Van Oostveldt P, Mikoshiba K, Diamond MP, Dhont M. 2002.
Inositol 1,4,5-trisphosphate receptor function in human oocytes: Calcium responses and oocyte
activation-related phenomena induced by photolytic release of InsP3 are blocked by a specific
antibody to the type I receptor. Mol Hum Reprod 8: 912–918.
Gross SD, Schwab MS, Taieb FE, Lewellyn AL, Qian YW, Maller JL. 2000. The critical role of the
MAP kinase pathway in meiosis II in Xenopus oocytes is mediated by p90RSK. Curr Biol 10: 430–
438.
Hake LE, Richter JD. 1994. CPEB is a specificity factor that mediates cytoplasmic polyadenylation
during Xenopus oocyte maturation. Cell 79: 617–627.
Han SJ, Chen R, Paronetto MP, Conti M. 2005. WEE1B is an oocyte-specific kinase involved in the
control of meiotic arrest in the mouse. Curr Biol 15: 1670–1676.
Hansen DV, Tung JJ, Jackson PK. 2006. CaMKII and polo-like kinase 1 sequentially phosphorylate the
cytostatic factor Emi2/XErp1 to trigger its destruction and meiotic exit. Proc Natl Acad Sci 103:
608–613.
Harada Y, Matsumoto T, Hirahara S, Nakashima A, Ueno S, Oda S, Miyazaki S, Iwao Y. 2007.
Characterization of a sperm factor for egg activation at fertilization of the newt Cynops
pyrrhogaster. Dev Biol 306: 797–808.
Hashimoto N, Watanabe N, Furuta Y, Tamemoto H, Sagata N, Yokoyama M, Okazaki K, Nagayoshi
M, Takeda N, Ikawa Y, et al. 1994. Parthenogenetic activation of oocytes in c-mos-deficient mice.
Nature 370: 68–71.
Haywood M, Spaliviero J, Jimemez M, King NJ, Handelsman DJ, Allan CM. 2003. Sertoli and germ
cell development in hypogonadal (hpg) mice expressing transgenic follicle-stimulating hormone
alone or in combination with testosterone. Endocrinology 144: 509–517.
Hermo L, Pelletier RM, Cyr DG, Smith CE. 2010. Surfing the wave, cycle, life history, and
genes/proteins expressed by testicular germ cells. Part 1: Background to spermatogenesis,
spermatogonia, and spermatocytes. Microsc Res Tech 73: 241–278.
Heytens E, Parrington J, Coward K, Young C, Lambrecht S, Yoon SY, Fissore RA, Hamer R, Deane
CM, Ruas M, et al. 2009. Reduced amounts and abnormal forms of phospholipase Cζ (PLCζ) in
spermatozoa from infertile men. Hum Reprod 24: 2417–2428.
Hinckley M, Vaccari S, Horner K, Chen R, Conti M. 2005. The G-protein-coupled receptors GPR3 and
GPR12 are involved in cAMP signaling and maintenance of meiotic arrest in rodent oocytes. Dev
Biol 287: 249–261.
Hodgman R, Tay J, Mendez R, Richter JD. 2001. CPEB phosphorylation and cytoplasmic
polyadenylation are catalyzed by the kinase IAK1/Eg2 in maturing mouse oocytes. Development
128: 2815–2822.
Hofmann MC. 2008. Gdnf signaling pathways within the mammalian spermatogonial stem cell niche.
Mol Cell Endocrinol 288: 95–103.
Holdcraft RW, Braun RE. 2004. Hormonal regulation of spermatogenesis. Int J Androl 27: 335–342.
Hu J, Shima H, Nakagawa H. 1999. Glial cell line-derived neurotropic factor stimulates sertoli cell
proliferation in the early postnatal period of rat testis development. Endocrinology 140: 3416–3421.
Hung PH, Suarez SS. 2010. Regulation of sperm storage and movement in the ruminant oviduct. Soc
Reprod Fertil Suppl 67: 257–266.
Hunter AG, Moor RM. 1987. Stage-dependent effects of inhibiting ribonucleic acids and protein
synthesis on meiotic maturation of bovine oocytes in vitro. J Dairy Sci 70: 1646–1651.
Iino M. 1990a. Biphasic Ca2+ dependence of inositol 1,4,5-trisphosphate-induced Ca release in smooth
muscle cells of the guinea pig taenia caeci. J Gen Physiol 95: 1103–1122.
Iino M. 1990b. Calcium release mechanisms in smooth muscle. Jpn J Pharmacol 54: 345–354.
Ikawa M, Inoue N, Benham AM, Okabe M. 2010. Fertilization: A sperm’s journey to and interaction
with the oocyte. J Clin Invest 120: 984–994.
Inoue N, Ikawa M, Isotani A, Okabe M. 2005. The immunoglobulin superfamily protein Izumo is
required for sperm to fuse with eggs. Nature 434: 234–238.
Inoue N, Yamaguchi R, Ikawa M, Okabe M. 2007. Sperm-egg interaction and gene manipulated
animals. Soc Reprod Fertil Suppl 65: 363–371.
Itman C, Mendis S, Barakat B, Loveland KL. 2006. All in the family: TGF-β family action in testis
development. Reproduction 132: 233–246.
Ito M, Shikano T, Oda S, Horiguchi T, Tanimoto S, Awaji T, Mitani H, Miyazaki S. 2008. Difference
in Ca2+ oscillation-inducing activity and nuclear translocation ability of PLCZ1, an egg-activating
sperm factor candidate, between mouse, rat, human, and medaka fish. Biol Reprod 78: 1081–1090.
Iwasaki H, Chiba K, Uchiyama T, Yoshikawa F, Suzuki F, Ikeda M, Furuichi T, Mikoshiba K. 2002.
Molecular characterization of the starfish inositol 1,4,5-trisphosphate receptor and its role during
oocyte maturation and fertilization. J Biol Chem 277: 2763–2772.
Jagiello GM. 1969. Meiosis and inhibition of ovulation in mouse eggs treated with actinomycin D. J
Cell Biol 42: 571–574.
Jellerette T, He CL, Wu H, Parys JB, Fissore RA. 2000. Down-regulation of the inositol 1,4,5-
trisphosphate receptor in mouse eggs following fertilization or parthenogenetic activation. Dev Biol
223: 238–250.
Jin M, Fujiwara E, Kakiuchi Y, Okabe M, Satouh Y, Baba SA, Chiba K, Hirohashi N. 2011. Most
fertilizing mouse spermatozoa begin their acrosome reaction before contact with the zona pellucida
during in vitro fertilization. Proc Natl Acad Sci 108: 4892–4896.
Jones KT, Matsuda M, Parrington J, Katan M, Swann K. 2000. Different Ca2+-releasing abilities of
sperm extracts compared with tissue extracts and phospholipase C isoforms in sea urchin egg
homogenate and mouse eggs. Biochem J 346: 743–749.
Kanki JP, Donoghue DJ. 1991. Progression from meiosis I to meiosis II in Xenopus oocytes requires de
novo translation of the mosxe protooncogene. Proc Natl Acad Sci 88: 5794–5798.
Katayama Y, Miyazaki S, Oshimi Y, Oshimi K. 1993. Ca2+ response in single human T cells induced
by stimulation of CD4 or CD8 and interference with CD3 stimulation. J Immunol Meth 166: 145–
153.
Kim E, Yamashita M, Kimura M, Honda A, Kashiwabara SI, Baba T. 2008. Sperm penetration through
cumulus mass and zona pellucida. Int J Dev Biol 52: 677–682.
Kim KS, Foster JA, Kvasnicka KW, Gerton GL. 2011. Transitional states of acrosomal exocytosis and
proteolytic processing of the acrosomal matrix in guinea pig sperm. Mol Reprod Dev 78: 930–941.
Kouchi Z, Fukami K, Shikano T, Oda S, Nakamura Y, Takenawa T, Miyazaki S. 2004. Recombinant
phospholipase Cζ has high Ca2+ sensitivity and induces Ca2+ oscillations in mouse eggs. J Biol
Chem 279: 10408–10412.
Kumagai A, Dunphy WG. 1999. Binding of 14-3-3 proteins and nuclear export control the intracellular
localization of the mitotic inducer Cdc25. Genes Dev 13: 1067–1072.
Kume S, Yamamoto A, Inoue T, Muto A, Okano H, Mikoshiba K. 1997. Developmental expression of
the inositol 1,4,5-trisphosphate receptor and structural changes in the endoplasmic reticulum during
oogenesis and meiotic maturation of Xenopus laevis. Dev Biol 182: 228–239.
Kurokawa M, Fissore RA. 2003. ICSI-generated mouse zygotes exhibit altered calcium oscillations,
inositol 1,4,5-trisphosphate receptor-1 down-regulation, and embryo development. Mol Hum
Reprod 9: 523–533.
Kyozuka K, Deguchi R, Mohri T, Miyazaki S. 1998. Injection of sperm extract mimics spatiotemporal
dynamics of Ca2+ responses and progression of meiosis at fertilization of ascidian oocytes.
Development 125: 4099–4105.
Labbe JC, Picard A, Peaucellier G, Cavadore JC, Nurse P, Doree M. 1989. Purification of MPF from
starfish: Identification as the H1 histone kinase p34cdc2 and a possible mechanism for its periodic
activation. Cell 57: 253–263.
Le Naour F, Rubinstein E, Jasmin C, Prenant M, Boucheix C. 2000. Severely reduced female fertility in
CD9-deficient mice. Science 287: 319–321.
Li YF, He W, Jha KN, Klotz K, Kim YH, Mandal A, Pulido S, Digilio L, Flickinger CJ, Herr JC. 2007.
FSCB, a novel protein kinase A-phosphorylated calcium-binding protein, is a CABYR-binding
partner involved in late steps of fibrous sheath biogenesis. J Biol Chem 282: 34104–34119.
Liou J, Kim ML, Heo WD, Jones JT, Myers JW, Ferrell JE Jr, Meyer T. 2005. STIM is a Ca2+ sensor
essential for Ca2+-store-depletion-triggered Ca2+ influx. Curr Biol 15: 1235–1241.
Lishko PV, Botchkina IL, Kirichok Y. 2011. Progesterone activates the principal Ca2+ channel of
human sperm. Nature 471: 387–391.
Liu J, Maller JL. 2005. Calcium elevation at fertilization coordinates phosphorylation of XErp1/Emi2
by Plx1 and CaMK II to release metaphase arrest by cytostatic factor. Curr Biol 15: 1458–1468.
Lopez-Girona A, Furnari B, Mondesert O, Russell P. 1999. Nuclear localization of Cdc25 is regulated
by DNA damage and a 14-3-3 protein. Nature 397: 172–175.
Lu Q, Shur BD. 1997. Sperm from β1,4-galactosyltransferase-null mice are refractory to ZP3-induced
acrosome reactions and penetrate the zona pellucida poorly. Development 124: 4121–4131.
Madgwick S, Hansen DV, Levasseur M, Jackson PK, Jones KT. 2006. Mouse Emi2 is required to enter
meiosis II by reestablishing cyclin B1 during interkinesis. J Cell Biol 174: 791–801.
Mahbub Hasan AK, Sato K, Sakakibara K, Ou Z, Iwasaki T, Ueda Y, Fukami Y. 2005. Uroplakin III, a
novel Src substrate in Xenopus egg rafts, is a target for sperm protease essential for fertilization.
Dev Biol 286: 483–492.
Malcuit C, Knott JG, He C, Wainwright T, Parys JB, Robl JM, Fissore RA. 2005. Fertilization and
inositol 1,4,5-trisphosphate (IP3)-induced calcium release in type-1 inositol 1,4,5-trisphosphate
receptor down-regulated bovine eggs. Biol Reprod 73: 2–13.
Malcuit C, Maserati M, Takahashi Y, Page R, Fissore RA. 2006. Intracytoplasmic sperm injection in
the bovine induces abnormal [Ca2+]i responses and oocyte activation. Reprod Fertil Dev 18: 39–
51.
Masui Y, Markert CL. 1971. Cytoplasmic control of nuclear behavior during meiotic maturation of frog
oocytes. J Exp Zool 177: 129–145.
Matthiesson KL, McLachlan RI, O’Donnell L, Frydenberg M, Robertson DM, Stanton PG, Meachem
SJ. 2006. The relative roles of follicle-stimulating hormone and luteinizing hormone in maintaining
spermatogonial maturation and spermiation in normal men. J Clin Endocrinol Metab 91: 3962–
3969.
Mattioli M, Galeati G, Bacci ML, Barboni B. 1991. Changes in maturation-promoting activity in the
cytoplasm of pig oocytes throughout maturation. Mol Reprod Dev 30: 119–125.
Mayorga LS, Tomes CN, Belmonte SA. 2007. Acrosomal exocytosis, a special type of regulated
secretion. IUBMB Life 59: 286–292.
McGuinness OM, Moreton RB, Johnson MH, Berridge MJ. 1996. A direct measurement of increased
divalent cation influx in fertilised mouse oocytes. Development 122: 2199–2206.
McLachlan RI, O’Donnell L, Meachem SJ, Stanton PG, de Kretser DM, Pratis K, Robertson DM.
2002. Identification of specific sites of hormonal regulation in spermatogenesis in rats, monkeys,
and man. Recent Prog Horm Res 57: 149–179.
Mehlmann LM. 2005. Oocyte-specific expression of Gpr3 is required for the maintenance of meiotic
arrest in mouse oocytes. Dev Biol 288: 397–404.
Mendez R, Hake LE, Andresson T, Littlepage LE, Ruderman JV, Richter JD. 2000. Phosphorylation of
CPE binding factor by Eg2 regulates translation of c-mos mRNA. Nature 404: 302–307.
Meng X, Lindahl M, Hyvonen ME, Parvinen M, de Rooij DG, Hess MW, Raatikainen-Ahokas A,
Sainio K, Rauvala H, Lakso M, et al. 2000. Regulation of cell fate decision of undifferentiated
spermatogonia by GDNF. Science 287: 1489–1493.
Miyado K, Yamada G, Yamada S, Hasuwa H, Nakamura Y, Ryu F, Suzuki K, Kosai K, Inoue K,
Ogura A, et al. 2000. Requirement of CD9 on the egg plasma membrane for fertilization. Science
287: 321–324.
Miyazaki S, Ito M. 2006. Calcium signals for egg activation in mammals. J Pharmacol Sci 100: 545–
552.
Miyazaki S, Yuzaki M, Nakada K, Shirakawa H, Nakanishi S, Nakade S, Mikoshiba K. 1992. Block of
Ca2+ wave and Ca2+ oscillation by antibody to the inositol 1,4,5-trisphosphate receptor in
fertilized hamster eggs. Science 257: 251–255.
Moor RM, Crosby IM. 1986. Protein requirements for germinal vesicle breakdown in ovine oocytes. J
Embryol Exp Morphol 94: 207–220.
Nakano Y, Shirakawa H, Mitsuhashi N, Kuwabara Y, Miyazaki S. 1997. Spatiotemporal dynamics of
intracellular calcium in the mouse egg injected with a spermatozoon. Mol Hum Reprod 3: 1087–
1093.
Nalam RL, Matzuk MM. 2010. Local signalling environments and human male infertility: What we can
learn from mouse models. Expert Rev Mol Med 12: e15.
Naughton CK, Jain S, Strickland AM, Gupta A, Milbrandt J. 2006. Glial cell-line derived neurotrophic
factor-mediated RET signaling regulates spermatogonial stem cell fate. Biol Reprod 74: 314–321.
Nishizawa M, Furuno N, Okazaki K, Tanaka H, Ogawa Y, Sagata N. 1993. Degradation of Mos by the
N-terminal proline (Pro2)-dependent ubiquitin pathway on fertilization of Xenopus eggs: Possible
significance of natural selection for Pro2 in Mos. EMBO J 12: 4021–4027.
Oatley JM, Brinster RL. 2012. The germline stem cell niche unit in mammalian testes. Physiol Rev 92:
577–595.
Ohe M, Inoue D, Kanemori Y, Sagata N. 2007. Erp1/Emi2 is essential for the meiosis I to meiosis II
transition in Xenopus oocytes. Dev Biol 303: 157–164.
Osman RA, Andria ML, Jones AD, Meizel S. 1989. Steroid induced exocytosis: The human sperm
acrosome reaction. Biochem Biophys Res Commun 160: 828–833.
Palmer A, Gavin AC, Nebreda AR. 1998. A link between MAP kinase and p34cdc2/cyclin B during
oocyte maturation: p90RSK phosphorylates and inactivates the p34cdc2 inhibitory kinase Myt1.
EMBO J 17: 5037–5047.
Park CY, Hoover PJ, Mullins FM, Bachhawat P, Covington ED, Raunser S, Walz T, Garcia KC,
Dolmetsch RE, Lewis RS. 2009. STIM1 clusters and activates CRAC channels via direct binding of
a cytosolic domain to Orai1. Cell 136: 876–890.
Parrington J, Jones KT, Lai A, Swann K. 1999. The soluble sperm factor that causes Ca2+ release from
sea-urchin (Lytechinus pictus) egg homogenates also triggers Ca2+ oscillations after injection into
mouse eggs. Biochem J 341: 1–4.
Parrington J, Davis LC, Galione A, Wessel G. 2007. Flipping the switch: How a sperm activates the
egg at fertilization. Dev Dyn 236: 2027–2038.
Parys JB, McPherson SM, Mathews L, Campbell KP, Longo FJ. 1994. Presence of inositol 1,4,5-
trisphosphate receptor, calreticulin, and calsequestrin in eggs of sea urchins and Xenopus laevis.
Dev Biol 161: 466–476.
Propst F, Rosenberg MP, Iyer A, Kaul K, Vande Woude GF. 1987. c-mos proto-oncogene RNA
transcripts in mouse tissues: Structural features, developmental regulation, and localization in
specific cell types. Mol Cell Biol 7: 1629–1637.
Rauh NR, Schmidt A, Bormann J, Nigg EA, Mayer TU. 2005. Calcium triggers exit from meiosis II by
targeting the APC/C inhibitor XErp1 for degradation. Nature 437: 1048–1052.
Ren D, Navarro B, Perez G, Jackson AC, Hsu S, Shi Q, Tilly JL, Clapham DE. 2001. A sperm ion
channel required for sperm motility and male fertility. Nature 413: 603–609.
* Rhind N, Russell P. 2012. Signaling pathways that regulate cell division. Cold Spring Harb Perspect
Biol 4: a005942.
Rice A, Parrington J, Jones KT, Swann K. 2000. Mammalian sperm contain a Ca2+-sensitive
phospholipase C activity that can generate InsP3 from PIP2 associated with intracellular organelles.
Dev Biol 228: 125–135.
Richards JS, Pangas SA. 2010. The ovary: Basic biology and clinical implications. J Clin Invest 120:
963–972.
Rogers NT, Hobson E, Pickering S, Lai FA, Braude P, Swann K. 2004. Phospholipase Czeta causes
Ca2+ oscillations and parthenogenetic activation of human oocytes. Reproduction 128: 697–702.
Roos J, DiGregorio PJ, Yeromin AV, Ohlsen K, Lioudyno M, Zhang S, Safrina O, Kozak JA, Wagner
SL, Cahalan MD, et al. 2005. STIM1, an essential and conserved component of store-operated
Ca2+ channel function. J Cell Biol 169: 435–445.
Ross PJ, Beyhan Z, Iager AE, Yoon SY, Malcuit C, Schellander K, Fissore RA, Cibelli JB. 2008.
Parthenogenetic activation of bovine oocytes using bovine and murine phospholipase Cζ. BMC Dev
Biol 8: 16.
Ruiz EJ, Hunt T, Nebreda AR. 2008. Meiotic inactivation of Xenopus Myt1 by CDK/XRINGO, but not
CDK/cyclin, via site-specific phosphorylation. Mol Cell 32: 210–220.
Runft LL, Watras J, Jaffe LA. 1999. Calcium release at fertilization of Xenopus eggs requires type I IP3
receptors, but not SH2 domain-mediated activation of PLCγ or Gq-mediated activation of PLCβ.
Dev Biol 214: 399–411.
Ruwanpura SM, McLachlan RI, Stanton PG, Meachem SJ. 2008. Follicle-stimulating hormone affects
spermatogonial survival by regulating the intrinsic apoptotic pathway in adult rats. Biol Reprod 78:
705–713.
Ruwanpura SM, McLachlan RI, Meachem SJ. 2010. Hormonal regulation of male germ cell
development. J Endocrinol 205: 117–131.
Sagata N, Watanabe N, Vande Woude GF, Ikawa Y. 1989. The c-mos proto-oncogene product is a
cytostatic factor responsible for meiotic arrest in vertebrate eggs. Nature 342: 512–518.
Saito K, O’Donnell L, McLachlan RI, Robertson DM. 2000. Spermiation failure is a major contributor
to early spermatogenic suppression caused by hormone withdrawal in adult rats. Endocrinology
141: 2779–2785.
Salicioni AM, Platt MD, Wertheimer EV, Arcelay E, Allaire A, Sosnik J, Visconti PE. 2007. Signalling
pathways involved in sperm capacitation. Soc Reprod Fertil Suppl 65: 245–259.
Sariola H, Saarma M. 2003. Novel functions and signalling pathways for GDNF. J Cell Sci 116: 3855–
3862.
Sato K, Tokmakov AA, Iwasaki T, Fukami Y. 2000. Tyrosine kinase-dependent activation of
phospholipase Cγ is required for calcium transient in Xenopus egg fertilization. Dev Biol 224: 453–
469.
Saunders CM, Larman MG, Parrington J, Cox LJ, Royse J, Blayney LM, Swann K, Lai FA. 2002.
PLCζ: A sperm-specific trigger of Ca2+ oscillations in eggs and embryo development.
Development 129: 3533–3544.
Schultz RM, Kopf GS. 1995. Molecular-basis of mammalian egg activation. Curr Top Dev Biol 30:
21–62.
Sette C, Dolci S, Geremia R, Rossi P. 2000. The role of stem cell factor and of alternative c-kit gene
products in the establishment, maintenance and function of germ cells. Int J Dev Biol 44: 599–608.
* Sever R, Glass CK. 2013. Signaling by nuclear receptors. Cold Spring Harb Perspect Biol 5:
a016709.
Shamsadin R, Adham IM, Nayernia K, Heinlein UA, Oberwinkler H, Engel W. 1999. Male mice
deficient for germ-cell cyritestin are infertile. Biol Reprod 61: 1445–1451.
Sharpe RM. 1994. Regulation of spermatogenesis. Raven Press, New York.
Shupe J, Cheng J, Puri P, Kostereva N, Walker WH. 2011. Regulation of Sertoli-germ cell adhesion
and sperm release by FSH and nonclassical testosterone signaling. Mol Endocrinol 25: 238–252.
Signorelli J, Diaz ES, Morales P. 2012. Kinases, phosphatases and proteases during sperm capacitation.
Cell Tissue Res 349: 765–782.
Simon L, Ekman GC, Tyagi G, Hess RA, Murphy KM, Cooke PS. 2007. Common and distinct factors
regulate expression of mRNA for ETV5 and GDNF, Sertoli cell proteins essential for
spermatogonial stem cell maintenance. Exp Cell Res 313: 3090–3099.
Smyth JT, Dehaven WI, Jones BF, Mercer JC, Trebak M, Vazquez G, Putney JW Jr. 2006. Emerging
perspectives in store-operated Ca2+ entry: Roles of Orai, Stim and TRP. Biochim Biophys Acta
1763: 1147–1160.
Sorrentino V, Giorgi M, Geremia R, Besmer P, Rossi P. 1991. Expression of the c-kit proto-oncogene
in the murine male germ cells. Oncogene 6: 149–151.
Sosnik J, Miranda PV, Spiridonov NA, Yoon SY, Fissore RA, Johnson GR, Visconti PE. 2009. Tssk6
is required for Izumo relocalization and gamete fusion in the mouse. J Cell Sci 122: 2741–2749.
Stanford JS, Ruderman JV. 2005. Changes in regulatory phosphorylation of Cdc25C Ser287 and WEE1
Ser549 during normal cell cycle progression and checkpoint arrests. Mol Biol Cell 16: 5749–5760.
Stanford JS, Lieberman SL, Wong VL, Ruderman JV. 2003. Regulation of the G2/M transition in
oocytes of Xenopus tropicalis. Dev Biol 260: 438–448.
Steegborn C, Litvin TN, Levin LR, Buck J, Wu H. 2005. Bicarbonate activation of adenylyl cyclase via
promotion of catalytic active site closure and metal recruitment. Nat Struct Mol Biol 12: 32–37.
Stice SL, Robl JM. 1990. Activation of mammalian oocytes by a factor obtained from rabbit sperm.
Mol Reprod Dev 25: 272–280.
Stricker SA. 1997. Intracellular injections of a soluble sperm factor trigger calcium oscillations and
meiotic maturation in unfertilized oocytes of a marine worm. Dev Biol 186: 185–201.
Stricker SA. 1999. Comparative biology of calcium signaling during fertilization and egg activation in
animals. Dev Biol 211: 157–176.
Stricker SA, Whitaker M. 1999. Confocal laser scanning microscopy of calcium dynamics in living
cells. Microsc Res Tech 46: 356–369.
Strunker T, Goodwin N, Brenker C, Kashikar ND, Weyand I, Seifert R, Kaupp UB. 2011. The CatSper
channel mediates progesterone-induced Ca2+ influx in human sperm. Nature 471: 382–386.
Suarez SS. 2008a. Control of hyperactivation in sperm. Hum Reprod Update 14: 647–657.
Suarez SS. 2008b. Regulation of sperm storage and movement in the mammalian oviduct. Int J Dev
Biol 52: 455–462.
Sun F, Bahat A, Gakamsky A, Girsh E, Katz N, Giojalas LC, Tur-Kaspa I, Eisenbach M. 2005. Human
sperm chemotaxis: Both the oocyte and its surrounding cumulus cells secrete sperm
chemoattractants. Hum Reprod 20: 761–767.
Swann K. 1990. A cytosolic sperm factor stimulates repetitive calcium increases and mimics
fertilization in hamster eggs. Development 110: 1295–1302.
Swann K, Lai FA. 1997. A novel signalling mechanism for generating Ca2+ oscillations at fertilization
in mammals. BioEssays 19: 371–378.
Swann K, Saunders CM, Rogers NT, Lai FA. 2006. PLCζ: A sperm protein that triggers Ca2+
oscillations and egg activation in mammals. Sem Cell Dev Biol 17: 264–273.
Tang W, Wu JQ, Guo Y, Hansen DV, Perry JA, Freel CD, Nutt L, Jackson PK, Kornbluth S. 2008.
Cdc2 and Mos regulate Emi2 stability to promote the meiosis I-meiosis II transition. Mol Biol Cell
19: 3536–3543.
Tesarik J, Testart J. 1994. Treatment of sperm-injected human oocytes with Ca2+ ionophore supports
the development of Ca2+ oscillations. Biol Reprod 51: 385–391.
Thomas TW, Eckberg WR, Dube F, Galione A. 1998. Mechanisms of calcium release and
sequestration in eggs of Chaetopterus pergamentaceus. Cell Calcium 24: 285–292.
Tokmakov AA, Sato KI, Iwasaki T, Fukami Y. 2002. Src kinase induces calcium release in Xenopus
egg extracts via PLCγ and IP3-dependent mechanism. Cell Calcium 32: 11–20.
Tosti E, Boni R. 2004. Electrical events during gamete maturation and fertilization in animals and
humans. Hum Reprod Update 10: 53–65.
Tsafriri A, Chun SY, Zhang R, Hsueh AJ, Conti M. 1996. Oocyte maturation involves
compartmentalization and opposing changes of cAMP levels in follicular somatic and germ cells:
Studies using selective phosphodiesterase inhibitors. Dev Biol 178: 393–402.
Tung JJ, Hansen DV, Ban KH, Loktev AV, Summers MK, Adler JR 3rd, Jackson PK. 2005. A role for
the anaphase-promoting complex inhibitor Emi2/XErp1, a homolog of early mitotic inhibitor 1, in
cytostatic factor arrest of Xenopus eggs. Proc Natl Acad Sci 102: 4318–4323.
Uhlenbrock K, Gassenhuber H, Kostenis E. 2002. Sphingosine 1-phosphate is a ligand of the human
gpr3, gpr6 and gpr12 family of constitutively active G protein-coupled receptors. Cell Signal 14:
941–953.
van den Ham R, van Dissel-Emiliani FM, van Pelt AM. 2003. Expression of the scaffolding subunit A
of protein phosphatase 2A during rat testicular development. Biol Reprod 68: 1369–1375.
Vernet N, Dennefeld C, Rochette-Egly C, Oulad-Abdelghani M, Chambon P, Ghyselinck NB, Mark M.
2006. Retinoic acid metabolism and signaling pathways in the adult and developing mouse testis.
Endocrinology 147: 96–110.
Vig M, Beck A, Billingsley JM, Lis A, Parvez S, Peinelt C, Koomoa DL, Soboloff J, Gill DL, Fleig A,
et al. 2006. CRACM1 multimers form the ion-selective pore of the CRAC channel. Curr Biol 16:
2073–2079.
Visconti PE. 2009. Understanding the molecular basis of sperm capacitation through kinase design.
Proc Natl Acad Sci 106: 667–668.
Visconti PE, Krapf D, de la Vega-Beltran JL, Acevedo JJ, Darszon A. 2011. Ion channels,
phosphorylation and mammalian sperm capacitation. Asian J Androl 13: 395–405.
Wang RS, Yeh S, Tzeng CR, Chang C. 2009. Androgen receptor roles in spermatogenesis and fertility:
Lessons from testicular cell-specific androgen receptor knockout mice. Endocr Rev 30: 119–132.
Wasserman WJ, Masui Y. 1975. Effects of cyclohexamide on a cytoplasmic factor initiating meiotic
naturation in Xenopus oocytes. Exp Cell Res 91: 381–388.
Watanabe N, Broome M, Hunter T. 1995. Regulation of the human WEE1Hu CDK tyrosine 15-kinase
during the cell cycle. EMBO J 14: 1878–1891.
Whitaker M. 2006. Calcium at fertilization and in early development. Physiol Rev 86: 25–88.
Wu JQ, Kornbluth S. 2008. Across the meiotic divide—CSF activity in the post-Emi2/XErp1 era. J
Cell Sci 121: 3509–3514.
Wu H, He CL, Fissore RA. 1997. Injection of a porcine sperm factor triggers calcium oscillations in
mouse oocytes and bovine eggs. Mol Reprod Dev 46: 176–189.
Wu JQ, Hansen DV, Guo Y, Wang MZ, Tang W, Freel CD, Tung JJ, Jackson PK, Kornbluth S. 2007a.
Control of Emi2 activity and stability through Mos-mediated recruitment of PP2A. Proc Natl Acad
Sci 104: 16564–16569.
Wu Q, Guo Y, Yamada A, Perry JA, Wang MZ, Araki M, Freel CD, Tung JJ, Tang W, Margolis SS, et
al. 2007b. A role for Cdc2- and PP2A-mediated regulation of Emi2 in the maintenance of CSF
arrest. Curr Biol 17: 213–224.
Xu Z, Kopf GS, Schultz RM. 1994. Involvement of inositol 1,4,5-trisphosphate-mediated Ca2+ release
in early and late events of mouse egg activation. Development 120: 1851–1859.
Yamashita M, Honda A, Ogura A, Kashiwabara S, Fukami K, Baba T. 2008. Reduced fertility of
mouse epididymal sperm lacking Prss21/Tesp5 is rescued by sperm exposure to uterine
microenvironment. Genes Cells 13: 1001–1013.
Yanagimachi R. 1970. The movement of golden hamster spermatozoa before and after capacitation. J
Reprod Fertil 23: 193–196.
Yanagimachi R. 1994. Mammalian fertilization. In The physiology of reproduction, 2nd ed. (ed. Knobil
E, Neill JD), Vol. 1, pp. 189–317. Raven, New York.
Yanagimachi R, Mahi CA. 1976. The sperm acrosome reaction and fertilization in the guinea-pig: A
study in vivo. J Reprod Fertil 46: 49–54.
Yang J, Winkler K, Yoshida M, Kornbluth S. 1999. Maintenance of G2 arrest in the Xenopus oocyte: A
role for 14-3-3-mediated inhibition of Cdc25 nuclear import. EMBO J 18: 2174–2183.
Yoneda A, Kashima M, Yoshida S, Terada K, Nakagawa S, Sakamoto A, Hayakawa K, Suzuki K,
Ueda J, Watanabe T. 2006. Molecular cloning, testicular postnatal expression, and oocyte-
activating potential of porcine phospholipase Cζ. Reproduction 132: 393–401.
Yoon SY, Jellerette T, Salicioni AM, Lee HC, Yoo MS, Coward K, Parrington J, Grow D, Cibelli JB,
Visconti PE, et al. 2008. Human sperm devoid of PLC, zeta 1 fail to induce Ca2+ release and are
unable to initiate the first step of embryo development. J Clin Invest 118: 3671–3681.
Yoshida M, Sensui N, Inoue T, Morisawa M, Mikoshiba K. 1998. Role of two series of Ca2+
oscillations in activation of ascidian eggs. Dev Biol 203: 122–133.
Yoshida S, Sukeno M, Nabeshima Y. 2007. A vasculature-associated niche for undifferentiated
spermatogonia in the mouse testis. Science 317: 1722–1726.
Yoshinaga K, Nishikawa S, Ogawa M, Hayashi S, Kunisada T, Fujimoto T, Nishikawa S. 1991. Role of
c-kit in mouse spermatogenesis: Identification of spermatogonia as a specific site of c-kit
expression and function. Development 113: 689–699.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a006064
CHAPTER 18
SUMMARY
Outline
1 Introduction
2 The unfolded protein response
3 Stress signaling by MAP kinases
4 Conclusions
References
1 INTRODUCTION
An important aspect of cellular physiology is the maintenance of
homeostasis. Evolutionarily conserved biochemical mechanisms play a key
role in this process. Thus, exposure to extra- or intracellular stress disrupts
cellular homeostasis and causes the engagement of signaling pathways that
serve to rebalance biochemical processes within the cell. One example is the
AMP-activated protein kinase signaling pathway, which responds to
increased AMP and ADP concentrations within the cell by dampening
anabolic pathways and promoting catabolic pathways that replenish the ATP
supply (Ch. 14 [Hardie 2012]). A second example is the cellular response to
DNA damage that engages the ataxia telangiectasia mutated (ATM) stress-
signaling pathway to induce growth arrest mediated by the p53 tumor
suppressor protein and promote DNA repair before reentry into the cell cycle
(Ch. 6 [Rhind and Russell 2012]). A third example is the regulation of
receptor ligand sensitivity to control the amplitude of signal transduction.
This type of stress response maintains, for example, the dynamic range of
vision following exposure to high- and low-intensity light sources (see Ch. 11
[Julius and Nathans 2012]). Similarly, signaling by the metabolic hormones
leptin and insulin is dynamically regulated by stress-signaling pathways to
control feeding behavior and biosynthetic processes (Ch. 14 [Hardie 2012]).
Such pathways are critical for normal cellular homeostasis and adaptive
changes in cell physiology that benefit the organism.
In addition to their contributions to normal physiology, stress-activated
signaling pathways play roles in establishing dysfunctional states associated
with stress exposure and the development of disease. Here, we focus on two
different mammalian stress-activated-signaling pathways to illustrate these
concepts: the unfolded protein response (UPR) and stress-activated MAP
kinase (MAPK) pathways. The UPR is engaged within the endoplasmic
reticulum (ER) during biosynthetic stress and leads to a coordinated
inhibition of general protein translation and specific up-regulation of ER
functional capacity. The pathways involved therefore serve to maintain
cellular homeostasis. Stress-activated MAPK pathways, in contrast, are
regulated by a diverse array of intra- and extracellular stresses, including
environmental physical/chemical changes and exposure to inflammatory
cytokines. These stress pathways cause phosphorylation of nuclear and
cytoplasmic substrates, leading to a network response and adaptation to the
new cellular environment.
The UPR and MAPK pathways can function separately or cooperatively.
For example, increased saturated fatty acid levels, caused by a high-fat diet,
induce both the UPR and stress-activated MAPK pathways. Together, these
pathways cause adaptation to the new diet by regulating insulin signaling,
blood glucose concentration, and obesity (Fig. 1).
Figure 1. Stress-signaling pathways activated in response to metabolic stress. Feeding mice a high-fat
diet causes metabolic stress that leads to the UPR and activation of stress-activated MAP kinases.
These signaling pathways result in an adaptive response associated with obesity and altered insulin
sensitivity.
Early studies of adaptive ER responses showed that the levels of two ER-
localized chaperones, 78- and 94-kDa glucose-regulated proteins (GRP78 and
GRP94), are increased on glucose starvation (Shiu et al. 1977) and protein N-
glycosylation inhibitors and calcium ionophores enhance their expression
(Welch et al. 1983; Resendez et al. 1985; Kim et al 1987). Sambrook and
colleagues subsequently observed increased levels of ER chaperones on
overexpression of mutant influenza virus hemagglutinin and were the first to
propose that malfolded proteins in the lumen of the ER are detected and
invoke a response (Kozutsumi et al. 1988). Subsequently, genetic screens in
yeast identified IRE1 as an ER-localized receptor-like kinase and
ribonuclease required for the ER-to-nucleus signaling that activates
chaperone expression under ER stress conditions (Nikawa and Yamashita
1992; Cox et al. 1993; Mori et al. 1993). Hac1 was shown to be the leucine-
zipper transcription factor that functions downstream from IRE1 to induce
transcription by binding to a defined DNA sequence, the UPR element
(UPRE) (Cox and Walter 1996; Nikawa et al. 1996) in promoters turned on
by the UPR in yeast. The equivalent of Hac1 in multicellular organisms is
XBP1 (Shen et al. 2001; Calfon et al. 2002).
An additional cis-acting element (ER stress element, ERSE) was later
identified in promoter regions of mammalian ER chaperones and led to the
description of another ER-resident leucine-zipper transcription factor, ATF6
(Yoshida et al. 1998). Together, ATF6 and XBP1 stimulate the expression of
a broad array of genes involved in protein folding, secretion, and degradation
to clear misfolded proteins from the ER (Walter and Ron 2011). Finally, the
third molecule activated during the UPR was identified as PERK, one of the
four known eIF2α kinases (Baird and Wek 2012; Donnelly et al. 2013). It is
involved in translational attenuation, temporarily halting arrival of new
proteins in the ER (Harding et al. 1999). These three branches are now
considered as mediators of the canonical UPR (Fig. 2).
In the canonical model, the intraluminal domains of these initiators (i.e.,
the amino termini of IRE1 and PERK, and the carboxyl terminus of ATF6)
are bound by the chaperone Grp78 (also called BiP) in the absence of stress
and rendered inactive (Bertolotti et al. 2000; Shen et al. 2002). Accumulation
of improperly folded proteins in the ER lumen results in the recruitment of
BiP away from these UPR sensors. Stripping off BiP allows for
oligomerization and activation of PERK and IRE1, and translocation of
ATF6 to the Golgi cisternae, which lead to a cascade of downstream
signaling events (Shamu and Walter 1996; Bertolotti et al. 2000). Recent
studies support the view that more complex luminal events underlie mounting
of the UPR. For example, IRE1 can form higher-order oligomers on
activation in vitro (Li et al. 2010) and directly interact with unfolded proteins
(Gardner and Walter 2011). Dynamic regulation of IRE1 by BiP may thus
adjust the magnitude of activation as opposed to providing an “on-or-off”
switch (Pincus et al. 2010). Hence, it is likely that stress responses emanating
from the ER are more complex than the canonical UPR model.
Activation of the ATF6 branch of the UPR requires translocation of ATF6
to the Golgi body and processing by the serine protease site-1 protease and
the metalloprotease site-2 protease to release an active transcription factor
(Chen et al. 2002). This branch also responds to signals other than BiP
sequestration—for example, the redox status of the ER. Active ATF6 moves
to the nucleus to stimulate the expression of genes containing the ERSE1,
ERSE2, UPRE, and cAMP-response elements (p. 99 [Sassone-Corsi 2012])
in their promoters (Yoshida et al. 1998). Genes required for ER-associated
degradation and the gene encoding the ER degradation-enhancing α-
mannosidase-like protein (EDEM) contain UPREs and, when induced,
facilitate clearance and degradation of misfolded proteins from the ER lumen
(Yoshida et al. 1998; Friedlander et al. 2000; Kokame et al. 2001).
The oldest branch of the UPR is mediated by IRE1, which is conserved
from yeast to humans (Patil and Walter 2001; Calfon et al. 2002). IRE1 has
two known isoforms, α and β, the latter being restricted primarily to the
intestine (Wang et al. 1998). IRE1 harbors two distinct catalytic activities: a
serine/threonine kinase for which the only known substrate is IRE1 itself, and
an endoribonuclease activity (Sidrauski and Walter 1997). The
endoribonuclease activity is activated on dimerization and
autophosphorylation and cleaves a 26-nucleotide intron from the XBP1
messenger RNA (mRNA), generating an mRNA whose translation produces
functional XBP1 (so-called XBP1s) (Shamu and Walter 1996; Sidrauski and
Walter 1997). XBP1s, alone or in conjunction with ATF6α, launches a
transcriptional program that induces many ER chaperones (including BiP),
proteins involved in ER biogenesis, and secretion (for example, EDEM,
ERdj4, protein disulfide isomerase [PDI], and other ER proteins) (Yoshida et
al. 2001, 2003; Lee et al. 2003). The endonuclease activity of IRE1 can also
degrade other mRNAs, preventing their translation and thereby providing an
additional way to reduce the translational burden and thus relieve ER stress
(Hollien and Weissman 2006). This mechanism has been termed regulated
IRE1-dependent degradation.
GTP-bound eIF2 is essential for loading of the initiator Met-tRNA
(tRNA) onto an mRNA-charged 40S ribosomal subunit for translation
initiation (Hinnebusch and Lorsch 2012). Phosphorylation of its GTP-binding
subunit, eIF2α, at S51 by PERK is another important aspect of the UPR. This
converts eIF2α into a competitive inhibitor of eIF2B (the GTP exchange
factor for eIF2α). This sequesters eIF2B and reduces the rate of regeneration
of the eIF2-GTP-tRNAiMet ternary complex, which, in turn, results in lower
rates of global protein synthesis, thereby reducing the ER workload (Shi et al.
1998; Harding et al. 1999). At least three other kinases can phosphorylate
eIF2α at S51: double-stranded RNA-dependent kinase (PKR), general control
nonderepressible 2, and heme-regulated inhibitor kinase (Baird and Wek
2012; Donnelly et al. 2013). The PERK branch of the UPR is also linked to
transcriptional regulation through several distinct mechanisms, which
increase the level of and/or activate the transcription factors ATF2, ATF4,
C/EBP (Harding et al. 2000; Ma et al. 2002; Ron and Walter 2007), NRF2
(Cullinan et al. 2003), and NF-κB (Jiang et al. 2003; Deng et al. 2004).
Generation of the protein products of the induced transcripts is achieved in
the context of general translational attenuation through features in the
mRNAs that permit their preferential translation. For example, the 5′-end of
the ATF4 transcript has two upstream open reading frames (uORFs) that
prevent translation under normal circumstances (Somers et al. 2013).
However, under stressed conditions, ribosome capacitation is delayed, the
uORFs are skipped, and functional ribosome complexes are assembled at the
bona fide start codon (Harding et al. 2000). The synthesis of functional ATF4
consequently activates the expression of genes involved in apoptosis, ER
redox control, glucose metabolism, and the relief of eIF2α inhibition
(Harding et al. 2000; Ma et al. 2002; Jiang et al. 2004).
Figure 4. Stress-activated MAPK-signaling pathways. The p38 MAP kinases are primarily activated by
the MAPKK isoforms MKK3 and MKK6, but a minor contribution of MKK4 can be detected. All p38
MAPK isoforms are activated by MKK3 and MKK6, although p38δ is activated by MKK3
significantly more potently than MKK6. The JNK group of MAPKs is activated by the MAPKK
isoforms MKK4 and MKK7.
The JNKs are encoded by three genes (JNK1, JNK2, and JNK3), which
are alternatively spliced to yield 10 different isoforms (Gupta et al. 1996).
JNK1 and JNK2 are ubiquitously expressed, but JNK3 is expressed primarily
in the brain (Davis 2000). Gene disruption studies in mice show that these
JNK isoforms can mediate different biological responses (Davis 2000). The
p38 MAPKs are encoded by four genes that can be divided into two
subgroups (p38α/β and p38γ/δ). These p38 isoforms show nonredundant
functions and different sensitivities to small molecule inhibitors (Cuenda and
Rousseau 2007; Cuadrado and Nebreda 2010).
The minimal consensus sequence for target protein phosphorylation by
MAPKs is -S/T-P-. However, the presence of this consensus motif is not
sufficient for phosphorylation by a MAPK, which frequently requires a
docking interaction between the MAPK and another region of the substrate
(Enslen and Davis 2001; Tanoue and Nishida 2003; Akella et al. 2008). Two
types of MAPK docking motifs have been identified in substrates: (1) the
FXFP motif, and (2) the D domain, comprising a hydrophobic motif (LXL)
plus a basic region (Bardwell and Thorner 1996; Jacobs et al. 1999; Enslen
and Davis 2001). Structural analysis of proteins docked to p38 MAPK
(Chang et al. 2002) and JNK3 (Heo et al. 2004) show extensive interactions
between the docked proteins and regions of the MAPK outside the active site.
The sites of D domain and FXFP interaction on MAPKs are different (Akella
et al. 2008). In addition to these conserved interactions, the carboxy-terminal
sequence of p38γ MAPK can dock directly to proteins that have PDZ
domains (e.g., α1 syntrophin, PSD95 [also known as SAP90], and DLG [also
known as SAP97]) to direct their phosphorylation (Hasegawa et al. 1999;
Hou et al. 2010).
Bioinformatic analyses and protein interaction screens (e.g., two-hybrid
assays) have identified many MAPK substrates. More recently, chemical
genetic methods and mass spectroscopy have enabled a more comprehensive
analysis of MAPK substrates in specific tissues (Allen et al. 2007; Carlson et
al. 2011). These include membrane, cytosolic, and nuclear proteins that
participate in many biological processes, and especially many transcription
factors, such as ATF2 (activated by JNK1/2/3 and p38α/β MAPK), Jun
(activated by JNK1/2/3), MEF2C (activated by p38α/β MAPK and ERK5),
and Elk1 (activated by JNK1/2/3, p38α/β MAPK, and ERK1/2) (Whitmarsh
and Davis 2000). Other MAPK targets include protein kinases
phosphorylated and activated by MAPKs, including eEF2K (activated by
p38γ/δ MAPK), MK2/3 (activated by p38α/β MAPK), MK5 (activated by
ERK3/4), MNK1/2 and MSK1/2 (activated by ERK1/2 and p38α/β MAPK),
and RSK1/2/3 (activated by ERK1/2) (Cargnello and Roux 2011). Protein
phosphatases are also MAPK targets, including nuclear DUSP1 (protected
against proteasomal degradation by ERK phosphorylation) and cytoplasmic
DUSP6 (proteasomal degradation is promoted by ERK phosphorylation)
(Caunt and Keyse 2013). A complete understanding of MAPK function will
require a systems-level approach to define the network of interactions that
mediate MAPK signaling (Ch. 4 [Azeloglu and Iyengar 2014]). Nevertheless,
we already have a good understanding of the roles of these kinases in stress
signaling from biochemical and genetic experiments in multiple organisms.
4 CONCLUSIONS
Stress-signaling pathways are evolutionarily conserved and play an important
role in the maintenance of homeostasis and adaptation to new cellular
microenvironments (Fig. 5). Key areas for future research include structural
studies of stress-response signaling proteins and integrated analysis of the
signaling network’s response to stress.
Figure 5. Integrated response to metabolic stress. A high-fat diet causes metabolic stress associated
with increased amounts of saturated free fatty acids, which engage the UPR and stress-activated
MAPKs. The UPR includes three different signaling pathways that are initiated by IRE1, PERK, and
ATF6. The stress-activated MAPK response is initiated by the MLK group of MAPKKKs and leads to
the activation of the JNK and p38 MAPKs. Cross talk between the UPR and stress-activated MAPK
signaling leads to an integrated adaptive response.
ACKNOWLEDGMENTS
Studies in the Hotamisligil laboratory are currently supported by grants from
the National Institutes of Health (NIH), the Juvenile Diabetes Research
Foundation, the American Diabetes Association, Servier and UCB
Pharmaceuticals, and the Simmons Fund. Special thanks to Scott
Widenmaier, Ling Yang, and Takahisa Nakamura for critical discussions,
thoughtful comments, and help in preparing the manuscript, Ana Paula
Arruda and Suneng Fu for help in illustrations, and Megan Washack and
Claudia Garcia Wagner for editorial assistance. Studies in the Davis
laboratory are currently supported by grants from the NIH, the American
Diabetes Association, and the Howard Hughes Medical Institute. Kathy
Gemme provided expert editorial assistance. We thank the students and
fellows who contributed to the studies in our groups over the years and to our
collaborators. We regret the inadvertent omission of references to important
work by our colleagues because of space limitations.
REFERENCES
*Reference is in this book.
Abe MK, Kahle KT, Saelzler MP, Orth K, Dixon JE, Rosner MR. 2001. ERK7 is an autoactivated
member of the MAPK family. J Biol Chem 276: 21272–21279.
Aguirre V, Uchida T, Yenush L, Davis R, White MF. 2000. The c-Jun NH2-terminal kinase promotes
insulin resistance during association with insulin receptor substrate-1 and phosphorylation of
Ser307. J Biol Chem 275: 9047–9054.
Akella R, Moon TM, Goldsmith EJ. 2008. Unique MAP kinase binding sites. Biochim Biophys Acta
1784: 48–55.
Allen JJ, Li M, Brinkworth CS, Paulson JL, Wang D, Hübner A, Chou W-H, Davis RJ, Burlingame
AL, Messing RO, et al. 2007. A semi-synthetic epitope for kinase substrates. Nat Methods 4: 511–
516.
* Alto NM, Orth K. 2012. Subversion of cell signaling by pathogens. Cold Spring Harb Perspect Biol
4: a006114.
Appenzeller-Herzog C, Hall MN. 2012. Bidirectional crosstalk between endoplasmic reticulum stress
and mTOR signaling. Trends Cell Biol 22: 274–282.
Arbour N, Naniche D, Homann D, Davis RJ, Flavell RA, Oldstone MB. 2002. c-Jun NH2-terminal
kinase (JNK)1 and JNK2 signaling pathways have divergent roles in CD8+ T cell-mediated
antiviral immunity. J Exp Med 195: 801–810.
Arias E, Cuervo A. 2011. Chaperone-mediated autophagy in protein quality control. Curr Opin Cell
Biol 23: 184–189.
* Azeloglu EU, Iyengar R. 2014. Signaling networks: Information flow, computation, and decision
making. Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a005934.
Babour A, Bicknell A, Tourtellotte J, Niwa M. 2010. A surveillance pathway monitors the fitness of the
endoplasmic reticulum to control its inheritance. Cell 142: 256–269.
Baird TD, Wek RC. 2012. Eukaryotic initiation factor 2 phosphorylation and translational control in
metabolism. Adv Nutr 3: 307–321.
Bardwell L, Thorner J. 1996. A conserved motif at the amino termini of MEKs might mediate high-
affinity interaction with the cognate MAPKs. Trends Biochem Sci 21: 373–374.
Bertolotti A, Zhang Y, Hendershot LM, Harding HP, Ron D. 2000. Dynamic interaction of BiP and ER
stress transducers in the unfolded-protein response. Nat Cell Biol 2: 326–332.
Bhattacharyya RP, Reményi A, Good MC, Bashor CJ, Falick AM, Lim WA. 2006. The Ste5 scaffold
allosterically modulates signaling output of the yeast mating pathway. Science 311: 822–826.
Boden G, Duan X, Homko C, Molina EJ, Song W, Perez O, Cheung P, Merali S. 2008. Increase in
endoplasmic reticulum stress-related proteins and genes in adipose tissue of obese, insulin-resistant
individuals. Diabetes 57: 2438–2444.
Boden G, Song W, Duan X, Cheung P, Kresge K, Barrero C, Merali S. 2011. Infusion of glucose and
lipids at physiological rates causes acute endoplasmic reticulum stress in rat liver. Obesity 19:
1366–1373.
Brancho D, Tanaka N, Jaeschke A, Ventura JJ, Kelkar N, Tanaka Y, Kyuuma M, Takeshita T, Flavell
RA, Davis RJ. 2003. Mechanism of p38 MAP kinase activation in vivo. Genes Dev 17: 1969–1978.
Bromati CR, Lellis-Santos C, Yamanaka TS, Nogueira TCA, Leonelli M, Caperuto LC, Gorjao R,
Leite AR, Anhe GF, Bordin S. 2011. UPR induces transient burst of apoptosis in islets of early
lactating rats through reduced AKT phosphorylation via ATF4/CHOP stimulation of TRB3
expression. Am J Physiol 300: R92–R100.
Calfon M, Zeng H, Urano F, Till JH, Hubbard SR, Harding HP, Clark SG, Ron D. 2002. IRE1 couples
endoplasmic reticulum load to secretory capacity by processing the XBP-1 mRNA. Nature 415:
92–96.
Canagarajah BJ, Khokhlatchev A, Cobb MH, Goldsmith EJ. 1997. Activation mechanism of the MAP
kinase ERK2 by dual phosphorylation. Cell 90: 859–869.
Cargnello M, Roux PP. 2011. Activation and function of the MAPKs and their substrates, the MAPK-
activated protein kinases. Microbiol Mol Biol Rev 75: 50–83.
Caricilli A, Picardi P, de Abreu L, Ueno M, Prada P, Ropelle E, Hirabara S, Vieira P, Camara N, Curi
R, et al. 2011. Gut microbiota is a key modulator of insulin resistance in TLR 2 knockout mice.
PLoS Biol 9: e1001212.
Carlson SM, Chouinard CR, Labadorf A, Lam CJ, Schmelzle K, Fraenkel E, White FM. 2011. Large-
scale discovery of ERK2 substrates identifies ERK-mediated transcriptional regulation by ETV3.
Sci Signal 4: rs11.
Caunt CJ, Keyse SM. 2013. Dual-specificity MAP kinase phosphatases (MKPs): Shaping the outcome
of MAP kinase signalling. FEBS J 280: 489–504.
Chan J, Cooney G, Biden T, Laybutt D. 2011. Differential regulation of adaptive and apoptotic
unfolded protein response signalling by cytokine-induced nitric oxide production in mouse
pancreatic β cells. Diabetologia 54: 1766–1776.
Chang CI, Xu BE, Akella R, Cobb MH, Goldsmith EJ. 2002. Crystal structures of MAP kinase p38
complexed to the docking sites on its nuclear substrate MEF2A and activator MKK3b. Mol Cell 9:
1241–1249.
Chen ZJ. 2012. Ubiquitination in signaling to and activation of IKK. Immunol Rev 246: 95–106.
Chen X, Shen J, Prywes R. 2002. The luminal domain of ATF6 senses endoplasmic reticulum (ER)
stress and causes translocation of ATF6 from the ER to the Golgi. J Biol Chem 277: 13045–13052.
Chi H, Lu B, Takekawa M, Davis RJ, Flavell RA. 2004. GADD45β/GADD45γ and MEKK4 comprise
a genetic pathway mediating STAT4-independent IFNγ production in T cells. EMBO J 23: 1576–
1586.
Clark A, Dean J, Tudor C, Saklatvala J. 2009. Post-transcriptional gene regulation by MAP kinases via
AU-rich elements. Front Biosci 14: 847–871.
Conze D, Krahl T, Kennedy N, Weiss L, Lumsden J, Hess P, Flavell RA, Le Gros G, Davis RJ, Rincon
M. 2002. c-Jun NH2-terminal kinase (JNK)1 and JNK2 have distinct roles in CD8+ T cell
activation. J Exp Med 195: 811–823.
Copps KD, White MF. 2012. Regulation of insulin sensitivity by serine/threonine phosphorylation of
insulin receptor substrate proteins IRS1 and IRS2. Diabetologia 55: 2565–2582.
Cox JS, Walter P. 1996. A novel mechanism for regulating activity of a transcription factor that
controls the unfolded protein response. Cell 87: 391–404.
Cox JS, Shamu CE, Walter P. 1993. Transcriptional induction of genes encoding endoplasmic
reticulum resident proteins requires a transmembrane protein kinase. Cell 73: 1197–1206.
Coyle SM, Gilbert WV, Doudna JA. 2009. Direct link between RACK1 function and localization at the
ribosome in vivo. Mol Cell Biol 29: 1626–1634.
Cretenet G, Le Clech M, Gachon F. 2010. Circadian clock-coordinated 12 hr period rhythmic activation
of the IRE1α pathway controls lipid metabolism in mouse liver. Cell Metab 11: 47–57.
Cuadrado A, Nebreda AR. 2010. Mechanisms and functions of p38 MAPK signalling. Biochem J 429:
403–417.
Cuenda A, Rousseau S. 2007. p38 MAP-kinases pathway regulation, function and role in human
diseases. Biochim Biophys Acta 1773: 1358–1375.
Cullinan SB, Diehl JA. 2006. Coordination of ER and oxidative stress signaling: The PERK/Nrf2
signaling pathway. Int J Biochem Cell Biol 38: 317–332.
Cullinan SB, Zhang D, Hannink M, Arvisais E, Kaufman RJ, Diehl JA. 2003. Nrf2 is a direct PERK
substrate and effector of PERK-dependent cell survival. Mol Cell Biol 23: 7198–7209.
Das M, Sabio G, Jiang F, Rincon M, Flavell RA, Davis RJ. 2009. Induction of hepatitis by JNK-
mediated expression of TNF-α. Cell 136: 249–260.
Das M, Garlick DS, Greiner DL, Davis RJ. 2011. The role of JNK in the development of hepatocellular
carcinoma. Genes Dev 25: 634–645.
Davis RJ. 2000. Signal transduction by the JNK group of MAP kinases. Cell 103: 239–252.
Delepine M, Nicolino M, Barrett T, Golamaully M, Lathrop GM, Julier C. 2000. EIF2AK3, encoding
translation initiation factor 2α kinase 3, is mutated in patients with Wolcott-Rallison syndrome. Nat
Genet 25: 406–409.
Deleris P, Trost M, Topisirovic I, Tanguay PL, Borden KL, Thibault P, Meloche S. 2011. Activation
loop phosphorylation of ERK3/ERK4 by group I p21-activated kinases (PAKs) defines a novel
PAK-ERK3/4-MAPK-activated protein kinase 5 signaling pathway. J Biol Chem 286: 6470–6478.
Deng J, Lu PD, Zhang Y, Scheuner D, Kaufman RJ, Sonenberg N, Harding HP, Ron D. 2004.
Translational repression mediates activation of nuclear factor κB by phosphorylated translation
initiation factor 2. Mol Cell Biol 24: 10161–10168.
Dong C, Yang DD, Tournier C, Whitmarsh AJ, Xu J, Davis RJ, Flavell RA. 2000. JNK is required for
effector T-cell function but not for T-cell activation. Nature 405: 91–94.
Dong C, Davis RJ, Flavell RA. 2002. MAP kinases in the immune response. Annu Rev Immunol 20:
55–72.
Donnelly N, Gorman AM, Gupta S, Samali A. 2013. The eIF2α kinases: Their structures and functions.
Cell Mol Life Sci 70: 1252–1256.
Duesbery NS, Webb CP, Leppla SH, Gordon VM, Klimpel KR, Copeland TD, Ahn NG, Oskarsson
MK, Fukasawa K, Paull KD, et al. 1998. Proteolytic inactivation of MAP-kinase-kinase by anthrax
lethal factor. Science 280: 734–737.
Eguchi K, Manabe I, Oishi-Tanaka Y, Ohsugi M, Kono N, Ogata F, Yagi N, Ohto U, Kimoto M,
Miyake K, et al. 2012. Saturated fatty acid and TLR signaling link β cell dysfunction and islet
inflammation. Cell Metab 15: 518–533.
Ellgaard L, Helenius A. 2003. Quality control in the endoplasmic reticulum. Nat Rev Mol Cell Biol 4:
181–191.
English A, Zurek N, Voeltz G. 2009. Peripheral ER structure and function. Curr Opin Cell Biol 21:
596–602.
Enslen H, Davis RJ. 2001. Regulation of MAP kinases by docking domains. Biol Cell 93: 5–14.
Fanger GR, Johnson NL, Johnson GL. 1997. MEK kinases are regulated by EGF and selectively
interact with Rac/Cdc42. EMBO J 16: 4961–4972.
Fleming Y, Armstrong CG, Morrice N, Paterson A, Goedert M, Cohen P. 2000. Synergistic activation
of stress-activated protein kinase 1/c-Jun N-terminal kinase (SAPK1/JNK) isoforms by mitogen-
activated protein kinase kinase 4 (MKK4) and MKK7. Biochem J 352: 145–154.
Fonseca SG, Fukuma M, Lipson KL, Nguyen LX, Allen JR, Oka Y, Urano F. 2005. WFS1 is a novel
component of the unfolded protein response and maintains homeostasis of the endoplasmic
reticulum in pancreatic β-cells. J Biol Chem 280: 39609–39615.
Fonseca SG, Ishigaki S, Oslo CM, Lu S, Lipson KL, Ghosh R, Hayashi E, Ishihara H, Oka Y, Permutt
MA, et al. 2010. Wolfram syndrom 1 gene negatively regulate ER stress signaling in rodent and
human cells. J Clin Invest 120: 744–755.
Friedlander R, Jarosch E, Urban J, Volkwein C, Sommer T. 2000. A regulatory link between ER-
associated protein degradation and the unfolded-protein response. Nat Cell Biol 2: 379–384.
Fu S, Watkins SM, Hotamisligil GS. 2012. The role of endoplasmic reticulum in hepatic lipid
homeostasis and stress signaling. Cell Metab 15: 623–634.
Gallo KA, Johnson GL. 2002. Mixed-lineage kinase control of JNK and p38 MAPK pathways. Nat Rev
Mol Cell Biol 3: 663–672.
Gardner B, Walter P. 2011. Unfolded proteins are Ire1-activating ligands that directly induce the
unfolded protein response. Science 333: 1891–1894.
Garrington TP, Ishizuka T, Papst PJ, Chayama K, Webb S, Yujiri T, Sun W, Sather S, Russell DM,
Gibson SB, et al. 2000. MEKK2 gene disruption causes loss of cytokine production in response to
IgE and c-Kit ligand stimulation of ES cell-derived mast cells. EMBO J 19: 5387–5395.
Gee H, Noh S, Tang B, Kim K, Lee M. 2011. Rescue of ΔF508-CFTR trafficking via a GRASP-
dependent unconventional secretion pathway. Cell 146: 746–760.
Gething MJ, Sambrook J. 1992. Protein folding in the cell. Nature 355: 33–45.
Gonzalez-Teran B, Cortes JR, Manieri E, Matesanz N, Verdugo A, Rodriguez ME, Gonzalez-
Rodriguez A, Valverde A, Martin P, Davis RJ, et al. 2013. Eukaryotic elongation factor 2 controls
TNF-α translation in LPS-induced hepatitis. J Clin Invest 123: 164–178.
Good M, Tang G, Singleton J, Remenyi A, Lim WA. 2009. The Ste5 scaffold directs mating signaling
by catalytically unlocking the Fus3 MAP kinase for activation. Cell 136: 1085–1097.
Good MC, Zalatan JG, Lim WA. 2011. Scaffold proteins: Hubs for controlling the flow of cellular
information. Science 332: 680–686.
Gotoh T, Mori M. 2006. Nitric oxide and endoplasmic reticulum stress. Arterioscler Thromb Vasc Biol
26: 1439–1446.
* Green DR, Llambi F. 2014. Cell death signaling. Cold Spring Harb Perspect Biol doi:
10.1101/cshperspect.a006080.
Gregor MF, Hotamisligil GS. 2011. Inflammatory mechanisms in obesity. Annu Rev Immunol 29: 415–
445.
Gregor MF, Yang L, Fabbrini E, Mohammed BS, Eagon JC, Hotamisligil GS, Klein S. 2009.
Endoplasmic reticulum stress is reduced in tissues of obese subjects after weight loss. Diabetes 58:
693–700.
Gupta S, Barrett T, Whitmarsh AJ, Cavanagh J, Sluss HK, Derijard B, Davis RJ. 1996. Selective
interaction of JNK protein kinase isoforms with transcription factors. EMBO J 15: 2760–2770.
Han MS, Jung DY, Morel C, Lakhani SA, Kim JK, Flavell RA, Davis RJ. 2013. JNK expression by
macrophages promotes obesity-induced insulin resistance and inflammation. Science 339: 218–222.
* Hardie DG. 2012. Organismal carbohydrate and lipid homeostasis. Cold Spring Harb Perspect Biol
4: a006031.
Harding HP, Zhang Y, Ron D. 1999. Protein translation and folding are coupled by an endoplasmic-
reticulum-resident kinase. Nature 397: 271–274.
Harding HP, Novoa I, Zhang Y, Zeng H, Wek R, Schapira M, Ron D. 2000. Regulated translation
initiation controls stress-induced gene expression in mammalian cells. Mol Cell 6: 1099–1108.
Harding HP, Zeng H, Zhang Y, Jungries R, Chung P, Plesken H, Sabatini DD, Ron D. 2001. Diabetes
mellitus and exocrine pancreatic dysfunction in Perk-/- mice reveals a role for translational control
in secretory cell survival. Mol Cell 7: 1153–1163.
Harding HP, Zhang Y, Zeng H, Novoa I, Lu PD, Calfon M, Sadri N, Yun C, Popko B, Paules R, et al.
2003. An integrated stress response regulates amino acid metabolism and resistance to oxidative
stress. Mol Cell 11: 619–633.
Hasegawa M, Cuenda A, Spillantini MG, Thomas GM, Buée-Scherrer V, Cohen P, Goedert M. 1999.
Stress-activated protein kinase-3 interacts with the PDZ domain of α1-syntrophin. A mechanism for
specific substrate recognition. J Biol Chem 274: 12626–12631.
Hatori M, Hirota T, Iitsuka M, Kurabayashi N, Haraguchi S, Kokame K, Sato R, Nakai A, Miyata T,
Tsutsui K, et al. 2011. Light-dependent and circadian clock-regulated activation of sterol regulatory
element-binding protein, X-box-binding protein 1, and heat shock factor pathways. Proc Natl Acad
Sci 108: 4864–4869.
Hayashi T, Rizzuto R, Hajnoczky G, Su TP. 2009. MAM: More than just a housekeeper. Trends Cell
Biol 19: 81–88.
Heazlewood CK, Cook MC, Eri R, Price GR, Tauro SB, Taupin D, Thornton DJ, Png CW, Crockford
TL, Cornall RJ, et al. 2008. Aberrant mucin assembly in mice causes endoplasmic reticulum stress
and spontaneous inflammation resembling ulcerative colitis. PLoS Med 5: e54.
Heo YS, Kim SK, Seo CI, Kim YK, Sung BJ, Lee HS, Lee JI, Park SY, Kim JH, Hwang KY, et al.
2004. Structural basis for the selective inhibition of JNK1 by the scaffolding protein JIP1 and
SP600125. EMBO J 23: 2185–2195.
Hinnebusch AG, Lorsch JR. 2012. The mechanism of eukaryotic translation initiation: New insights
and challenges. Cold Spring Harb Perspect Biol 4: a011544.
Hirosumi J, Tuncman G, Chang L, Gorgun CZ, Uysal KT, Maeda K, Karin M, Hotamisligil GS. 2002.
A central role for JNK in obesity and insulin resistance. Nature 420: 333–336.
Holcik M, Sonenberg N. 2005. Translational control in stress and apoptosis. Nat Rev Mol Cell Biol 6:
318–327.
Hollien J, Weissman JS. 2006. Decay of endoplasmic reticulum-localized mRNAs during the unfolded
protein response. Science 313: 104–107.
Holzer RG, Park EJ, Li N, Tran H, Chen M, Choi C, Solinas G, Karin M. 2011. Saturated fatty acids
induce c-Src clustering within membrane subdomains, leading to JNK activation. Cell 147: 173–
184.
Hotamisligil GS. 2010. Endoplasmic reticulum stress and the inflammatory basis of metabolic disease.
Cell 140: 900–917.
Hou SW, Zhi HY, Pohl N, Loesch M, Qi XM, Li RS, Basir Z, Chen G. 2010. PTPH1 dephosphorylates
and cooperates with p38γ MAPK to increase ras oncogenesis through PDZ-mediated interaction.
Cancer Res 70: 2901–2910.
Hu P, Han Z, Couvillon AD, Kaufman RJ, Exton JH. 2006. Autocrine tumor necrosis factor α links
endoplasmic reticulum stress to the membrane death receptor pathway through IRE1α-mediated
NF-κB activation and down-regulation of TRAF2 expression. Mol Cell Biol 26: 3071–3084.
Hubner A, Barrett T, Flavell RA, Davis RJ. 2008. Multisite phosphorylation regulates Bim stability and
apoptotic activity. Mol Cell 30: 415–425.
Hubner A, Cavanagh-Kyros J, Rincon M, Flavell RA, Davis RJ. 2010. Functional cooperation of the
proapoptotic Bcl2 family proteins Bmf and Bim in vivo. Mol Cell Biol 30: 98–105.
Ishihara H, Takeda S, Tamura A, Takahashi R, Yamaguchi S, Takei D, Yamada T, Inoue H, Soga H,
Katagiri H, et al. 2004. Disruption of the WFS1 gene in mice causes progressive β-cell loss and
impaired stimulus-secretion coupling in insulin secretion. Hum Mol Genet 13: 1159–1170.
Jacobs D, Glossip D, Xing H, Muslin AJ, Kornfeld K. 1999. Multiple docking sites on substrate
proteins form a modular system that mediates recognition by ERK MAP kinase. Genes Dev 13:
163–175.
Jaeschke A, Davis RJ. 2007. Metabolic stress signaling mediated by mixed-lineage kinases. Mol Cell
27: 498–508.
Jaeschke A, Czech MP, Davis RJ. 2004. An essential role of the JIP1 scaffold protein for JNK
activation in adipose tissue. Genes Dev 18: 1976–1980.
Jiang HY, Wek SA, McGrath BC, Scheuner D, Kaufman RJ, Cavener DR, Wek RC. 2003.
Phosphorylation of the α subunit of eukaryotic initiation factor 2 is required for activation of NF-κB
in response to diverse cellular stresses. Mol Cell Biol 23: 5651–5663.
Jiang HY, Wek SA, McGrath BC, Lu D, Hai T, Harding HP, Wang X, Ron D, Cavener DR, Wek RC.
2004. Activating transcription factor 3 is integral to the eukaryotic initiation factor 2 kinase stress
response. Mol Cell Biol 24: 1365–1377.
Jiao P, Ma J, Feng B, Zhang H, Diehl JA, Chin YE, Yan W, Xu H. 2011. FFA-induced adipocyte
inflammation and insulin resistance: Involvement of ER stress and IKKβ pathways. Obesity 19:
483–491.
* Julius D, Nathans J. 2012. Signaling by sensory receptors. Cold Spring Harb Perspect Biol 4:
a005991.
Kammoun HL, Chabanon H, Hainault I, Luquet S, Magnan C, Koike T, Ferre P, Foufelle F. 2009.
GRP78 expression inhibits insulin and ER stress-induced SREBP-1c activation and reduces hepatic
steatosis in mice. J Clin Invest 119: 1201–1215.
Kaneko M, Niinuma Y, Nomura Y. 2003. Activation signal of nuclear factor-κB in response to
endoplasmic reticulum stress is transduced via IRE1 and tumor necrosis factor receptor-associated
factor 2. Biol Pharm Bull 26: 931–935.
Kant S, Swat W, Zhang S, Zhang ZY, Neel BG, Flavell RA, Davis RJ. 2011. TNF-stimulated MAP
kinase activation mediated by a Rho family GTPase signaling pathway. Genes Dev 25: 2069–2078.
Kant S, Barrett T, Vertii A, Noh YH, Jung DY, Kim JK, Davis RJ. 2013. Role of the mixed-lineage
protein kinase pathway in the metabolic stress response to obesity. Cell Rep 4: 681–688.
Kars M, Yang L, Gregor MF, Mohammed BS, Pietka TA, Finck BN, Patterson BW, Horton JD,
Mittendorfer B, Hotamisligil GS, et al. 2010. Tauroursodeoxycholic Acid may improve liver and
muscle but not adipose tissue insulin sensitivity in obese men and women. Diabetes 59: 1899–1905.
Kesavan K, Lobel-Rice K, Sun W, Lapadat R, Webb S, Johnson GL, Garrington TP. 2004. MEKK2
regulates the coordinate activation of ERK5 and JNK in response to FGF-2 in fibroblasts. J Cell
Physiol 199: 140–148.
Keyse SM. 2008. Dual-specificity MAP kinase phosphatases (MKPs) and cancer. Cancer Metastasis
Rev 27: 253–261.
Kim YK, Kim KS, Lee AS. 1987. Regulation of the glucose-regulated protein genes by β-
mercaptoethanol requires de novo protein synthesis and correlates with inhibition of protein
glycosylation. J Cell Physiol 133: 553–559.
Klionsky D. 2010. The molecular machinery of autophagy and its role in physiology and disease. Semin
Cell Dev Biol 21: 663.
Kokame K, Kato H, Miyata T. 2001. Identification of ERSE-II, a new cis-acting element responsible
for the ATF6-dependent mammalian unfolded protein response. J Biol Chem 276: 9199–9205.
Kostova Z, Wolf DH. 2003. For whom the bell tolls: Protein quality control of the endoplasmic
reticulum and the ubiquitin-proteasome connection. EMBO J 22: 2309–2317.
Kozutsumi Y, Segal M, Normington K, Gething MJ, Sambrook J. 1988. The presence of malfolded
proteins in the endoplasmic reticulum signals the induction of glucose-regulated proteins. Nature
332: 462–464.
Kroemer G, Marino G, Levine B. 2010. Autophagy and the integrated stress response. Mol Cell 40:
280–293.
Kuan CY, Whitmarsh AJ, Yang DD, Liao G, Schloemer AJ, Dong C, Bao J, Banasiak KJ, Haddad GG,
Flavell RA, et al. 2003. A critical role of neural-specific JNK3 for ischemic apoptosis. Proc Natl
Acad Sci 100: 15184–15189.
Ladiges WC, Knoblaugh SE, Morton JF, Korth MJ, Sopher BL, Baskin CR, MacAuley A, Goodman
AG, LeBoeuf RC, Katze MG. 2005. Pancreatic β-cell failure and diabetes in mice with a deletion
mutation of the endoplasmic reticulum molecular chaperone gene P58IPK. Diabetes 54: 1074–
1081.
Lamb JA, Ventura JJ, Hess P, Flavell RA, Davis RJ. 2003. JunD mediates survival signaling by the
JNK signal transduction pathway. Mol Cell 11: 1479–1489.
Lamkanfi M, Dixit VM. 2012. Inflammasomes and their roles in health and disease. Annu Rev Cell Dev
Biol 28: 137–161.
* Laplante M, Sabatini DM. 2012. mTOR signaling. Cold Spring Harb Perspect Biol 4: a011593.
Lee AH, Iwakoshi NN, Glimcher LH. 2003. XBP-1 regulates a subset of endoplasmic reticulum
resident chaperone genes in the unfolded protein response. Mol Cell Biol 23: 7448–7459.
Lee AH, Heidtman K, Hotamisligil GS, Glimcher LH. 2011a. Dual and opposing roles of the unfolded
protein response regulated by IRE1α and XBP1 in proinsulin processing and insulin secretion. Proc
Natl Acad Sci 108: 8885–8890.
Lee J, Sun C, Zhou Y, Lee J, Gokalp D, Herrema H, Park SW, Davis RJ, Ozcan U. 2011b. p38 MAPK-
mediated regulation of Xbp1s is crucial for glucose homeostasis. Nat Med 17: 1251–1260.
Le Guezennec X, Bulavin DV. 2010. WIP1 phosphatase at the crossroads of cancer and aging. Trends
Biochem Sci 35: 109–114.
Lei K, Davis RJ. 2003. JNK phosphorylation of Bim-related members of the Bcl2 family induces Bax-
dependent apoptosis. Proc Natl Acad Sci 100: 2432–2437.
Lerner AG, Upton JP, Praveen PV, Ghosh R, Nakagawa Y, Igbaria A, Shen S, Nguyen V, Backes BJ,
Heiman M, et al. 2012. IRE1α induces thioredoxin-interacting protein to activate the NLRP3
inflammasome and promote programmed cell death under irremediable ER stress. Cell Metab 16:
250–264.
Levine B, Mizushima N, Virgin H. 2011. Autophagy in immunity and inflammation. Nature 469: 323–
335.
Li Y, Schwabe RF, DeVries-Seimon T, Yao PM, Gerbod-Giannone MC, Tall AR, Davis RJ, Flavell R,
Brenner DA, Tabas I. 2005. Free cholesterol-loaded macrophages are an abundant source of tumor
necrosis factor-α and interleukin-6: Model of NF-κB- and map kinase-dependent inflammation in
advanced atherosclerosis. J Biol Chem 280: 21763–21772.
Li G, Mongillo M, Chin KT, Harding H, Ron D, Marks AR, Tabas I. 2009. Role of ERO1-α-mediated
stimulation of inositol 1,4,5-triphosphate receptor activity in endoplasmic reticulum stress-induced
apoptosis. J Cell Biol 186: 783–792.
Li H, Korennykh A, Behrman S, Walter P. 2010. Mammalian endoplasmic reticulum stress sensor
IRE1 signals by dynamic clustering. Proc Natl Acad Sci 107: 16113–16118.
* Lim K-H, Staudt LM. 2013. Toll-like receptor signaling. Cold Spring Harb Perspect Biol 5: a011247.
Lu B, Nakamura T, Inouye K, Li J, Tang Y, Lundback P, Valdes-Ferrer SI, Olofsson PS, Kalb T, Roth
J, et al. 2012. Novel role of PKR in inflammasome activation and HMGB1 release. Nature 488:
670–674.
Lynes E, Simmen T. 2011. Urban planning of the endoplasmic reticulum (ER): How diverse
mechanisms segregate the many functions of the ER. Biochim Biophys Acta 1813: 1893–1905.
Ma Y, Brewer JW, Diehl JA, Hendershot LM. 2002. Two distinct stress signaling pathways converge
upon the CHOP promoter during the mammalian unfolded protein response. J Mol Biol 318: 1351–
1365.
Marciniak SJ, Ron D. 2006. Endoplasmic reticulum stress signaling in disease. Physiol Rev 86: 1133–
1149.
Marciniak SJ, Yun CY, Oyadomari S, Novoa I, Zhang Y, Jungreis R, Nagata K, Harding HP, Ron D.
2004. CHOP induces death by promoting protein synthesis and oxidation in the stressed
endoplasmic reticulum. Genes Dev 18: 3066–3077.
Martinon F, Chen X, Lee A, Glimcher L. 2010. TLR activation of the transcription factor XBP1
regulates innate immune responses in macrophages. Nat Immunol 11: 411–418.
Matsukawa J, Matsuzawa A, Takeda K, Ichijo H. 2004. The ASK1-MAP kinase cascades in
mammalian stress response. J Biochem 136: 261–265.
Matsuzawa A, Nishitoh H, Tobiume K, Takeda K, Ichijo H. 2002. Physiological roles of ASK1-
mediated signal transduction in oxidative stress- and endoplasmic reticulum stress-induced
apoptosis: Advanced findings from ASK1 knockout mice. Antioxid Redox Signal 4: 415–425.
Mazzitelli S, Xu P, Ferrer I, Davis RJ, Tournier C. 2011. The loss of c-Jun N-terminal protein kinase
activity prevents the amyloidogenic cleavage of amyloid precursor protein and the formation of
amyloid plaques in vivo. J Neurosci 31: 16969–16976.
McAlees JW, Sanders VM. 2009. Hematopoietic protein tyrosine phosphatase mediates β2-adrenergic
receptor-induced regulation of p38 mitogen-activated protein kinase in B lymphocytes. Mol Cell
Biol 29: 675–686.
McCullough KD, Martindale JL, Klotz LO, Aw TY, Holbrook NJ. 2001. Gadd153 sensitizes cells to
endoplasmic reticulum stress by down-regulating Bcl2 and perturbing the cellular redox state. Mol
Cell Biol 21: 1249–1259.
McGuckin MA, Eri RD, Das I, Lourie R, Florin TH. 2011. Intestinal secretory cell ER stress and
inflammation. Biochem Soc Trans 39: 1081–1085.
Meusser B, Hirsch C, Jarosch E, Sommer T. 2005. ERAD: The long road to destruction. Nat Cell Biol
7: 766–772.
Mighiu PI, Filippi BM, Lam TK. 2012. Linking inflammation to the brain-liver axis. Diabetes 61:
1350–1352.
Milanski M, Arruda AP, Coope A, Ignacio-Souza LM, Nunez CE, Roman EA, Romanatto T, Pascoal
LB, Caricilli AM, Torsoni MA, et al. 2012. Inhibition of hypothalamic inflammation reverses diet-
induced insulin resistance in the liver. Diabetes 61: 1455–1462.
Miyake Z, Takekawa M, Ge Q, Saito H. 2007. Activation of MTK1/MEKK4 by GADD45 through
induced N-C dissociation and dimerization-mediated trans autophosphorylation of the MTK1
kinase domain. Mol Cell Biol 27: 2765–2776.
Morel C, Carlson SM, White FM, Davis RJ. 2009. Mcl-1 integrates the opposing actions of signaling
pathways that mediate survival and apoptosis. Mol Cell Biol 29: 3845–3852.
Morel C, Standen CL, Jung DY, Gray S, Ong H, Flavell RA, Kim JK, Davis RJ. 2010. Requirement of
JIP1-mediated c-Jun N-terminal kinase activation for obesity-induced insulin resistance. Mol Cell
Biol 30: 4616–4625.
Mori K. 2000. Tripartite management of unfolded proteins in the endoplasmic reticulum. Cell 101:
451–454.
Mori K, Ma W, Gething MJ, Sambrook J. 1993. A transmembrane protein with a cdc2+/CDC28-related
kinase activity is required for signaling from the ER to the nucleus. Cell 74: 743–756.
* Morrison DK. 2012. MAP kinase pathways. Cold Spring Harb Perspect Biol 4: a011254.
Morrison DK, Davis RJ. 2003. Regulation of MAP kinase signaling modules by scaffold proteins in
mammals. Annu Rev Cell Dev Biol 19: 91–118.
Nakamura K, Johnson GL. 2007. Noncanonical function of MEKK2 and MEK5 PB1 domains for
coordinated extracellular signal-regulated kinase 5 and c-Jun N-terminal kinase signaling. Mol Cell
Biol 27: 4566–4577.
Nakamura T, Furuhashi M, Li P, Cao H, Tuncman G, Sonenberg N, Gorgun CZ, Hotamisligil GS.
2010. Double-stranded RNA-dependent protein kinase links pathogen sensing with stress and
metabolic homeostasis. Cell 140: 338–348.
Nakatani Y, Kaneto H, Kawamori D, Yoshiuchi K, Hatazaki M, Matsuoka TA, Ozawa K, Ogawa S,
Hori M, Yamasaki Y, et al. 2005. Involvement of endoplasmic reticulum stress in insulin resistance
and diabetes. J Biol Chem 280: 847–851.
* Newton K, Dixit VM. 2012. Signaling in innate immunity and inflammation. Cold Spring Harb
Perspect Biol 4: a006049.
Niehrs C, Schäfer A. 2012. Active DNA demethylation by Gadd45 and DNA repair. Trends Cell Biol
22: 220–227.
Nikawa J, Yamashita S. 1992. IRE1 encodes a putative protein kinase containing a membrane-spanning
domain and is required for inositol phototrophy in Saccharomyces cerevisiae. Mol Microbiol 6:
1441–1446.
Nikawa J, Akiyoshi M, Hirata S, Fukuda T. 1996. Saccharomyces cerevisiae IRE2/HAC1 is involved
in IRE1-mediated KAR2 expression. Nucleic Acids Res 24: 4222–4226.
Noubade R, Krementsov DN, Del Rio R, Thornton T, Nagaleekar V, Saligrama N, Spitzack A, Spach
K, Sabio G, Davis RJ, et al. 2011. Activation of p38 MAPK in CD4 T cells controls IL-17
production and autoimmune encephalomyelitis. Blood 118: 3290–3300.
Novoa I, Zeng H, Harding HP, Ron D. 2001 Feedback inhibition of the unfolded protein response by
GADD34-mediated dephosphorylation of eIF2α. J Cell Biol 153: 1011–1022.
Nozaki J, Kubota H, Yoshida H, Naitoh M, Goji J, Yoshinaga T, Mori K, Koizumi A, Nagata K. 2004.
The endoplasmic reticulum stress response is stimulated through the continuous activation of
transcription factors ATF6 and XBP1 in Ins2+/Akita pancreatic β cells. Genes Cells 9: 261–270.
Ogata M, Hino S, Saito A, Morikawa K, Kondo S, Kanemoto S, Murakami T, Taniguchi M, Tanii I,
Yoshinaga K, et al. 2006. Autophagy is activated for cell survival after endoplasmic reticulum
stress. Mol Cell Biol 26: 9220–9231.
Oslowski CM, Hara T, O’Sullivan-Murphy B, Kanekura K, Lu S, Hara M, Ishigaki S, Zhu LJ, Hayashi
E, Hui ST, et al. 2012. Thioredoxin-interacting protein mediates ER stress-induced β cell death
through initiation of the inflammasome. Cell Metab 16: 265–273.
Owens DM, Keyse SM. 2007. Differential regulation of MAP kinase signalling by dual-specificity
protein phosphatases. Oncogene 26: 3203–3213.
Ozawa K, Miyazaki M, Matsuhisa M, Takano K, Nakatani Y, Hatazaki M, Tamatani T, Yamagata K,
Miyagawa J, Kitao Y, et al. 2005. The endoplasmic reticulum chaperone improves insulin
resistance in type 2 diabetes. Diabetes 54: 657–663.
Ozcan L, Tabas I. 2012. Role of endoplasmic reticulum stress in metabolic disease and other disorders.
Annu Rev Med 63: 317–328.
Ozcan U, Cao Q, Yilmaz E, Lee AH, Iwakoshi NN, Ozdelen E, Tuncman G, Gorgun C, Glimcher LH,
Hotamisligil GS. 2004. Endoplasmic reticulum stress links obesity, insulin action, and type 2
diabetes. Science 306: 457–461.
Ozcan U, Yilmaz E, Ozcan L, Furuhashi M, Vaillancourt E, Smith RO, Gorgun CZ, Hotamisligil GS.
2006. Chemical chaperones reduce ER stress and restore glucose homeostasis in a mouse model of
type 2 diabetes. Science 313: 1137–1140.
Ozcan U, Ozcan L, Yilmaz E, Duvel K, Sahin M, Manning BD, Hotamisligil GS. 2008. Loss of the
tuberous sclerosis complex tumor suppressors triggers the unfolded protein response to regulate
insulin signaling and apoptosis. Mol Cell 29: 541–551.
Ozcan L, Wong CC, Li G, Xu T, Pajvani U, Park SK, Wronska A, Chen BX, Marks AR, Fukamizu A,
et al. 2012. Calcium signaling through CaMKII regulates hepatic glucose production in fasting and
obesity. Cell Metab 15: 739–751.
Palade GE, Porter KR. 1954. Studies on the endoplasmic reticulum: I. Its identification in cells in situ. J
Exp Med 100: 641–656.
Patil C, Walter P. 2001. Intracellular signaling from the endoplasmic reticulum to the nucleus: The
unfolded protein response in yeast and mammals. Curr Opin Cell Biol 13: 349–355.
Payne DM, Rossomando AJ, Martino P, Erickson AK, Her JH, Shabanowitz J, Hunt DF, Weber MJ,
Sturgill TW. 1991. Identification of the regulatory phosphorylation sites in pp42/mitogen-activated
protein kinase (MAP kinase). EMBO J 10: 885–892.
Perier C, Bove J, Wu DC, Dehay B, Choi DK, Jackson-Lewis V, Rathke-Hartlieb S, Bouillet P,
Strasser A, Schulz JB, et al. 2007. Two molecular pathways initiate mitochondria-dependent
dopaminergic neurodegeneration in experimental Parkinson’s disease. Proc Natl Acad Sci 104:
8161–8166.
Pfaffenbach KT, Nivala AM, Reese L, Ellis F, Wang D, Wei Y, Pagliassotti MJ. 2010. Rapamycin
inhibits postprandial-mediated X-box-binding protein-1 splicing in rat liver. J Nutr 140: 879–884.
Pincus D, Chevalier M, Aragón T, Van Anken E, Vidal S, El-Samad H, Walter P. 2010. BiP binding to
the ER-stress sensor Ire1 tunes the homeostatic behavior of the unfolded protein response. PLoS
Biol 8: e1000415.
Proud CG. 2009. mTORC1 signalling and mRNA translation. Biochem Soc Trans 37: 227–231.
Purkayastha S, Zhang H, Zhang G, Ahmed Z, Wang Y, Cai D. 2011. Neural dysregulation of peripheral
insulin action and blood pressure by brain endoplasmic reticulum stress. Proc Natl Acad Sci 108:
2939–2944.
Puthalakath H, Strasser A. 2002. Keeping killers on a tight leash: Transcriptional and post-translational
control of the pro-apoptotic activity of BH3-only proteins. Cell Death Differ 9: 505–512.
Qiu Y, Mao T, Zhang Y, Shao M, You J, Ding Q, Chen Y, Wu D, Xie D, Lin X, et al. 2010. A crucial
role for RACK1 in the regulation of glucose-stimulated IRE1α activation in pancreatic β cells. Sci
Signal 3: ra7.
Raman M, Earnest S, Zhang K, Zhao Y, Cobb MH. 2007. TAO kinases mediate activation of p38 in
response to DNA damage. EMBO J 26: 2005–2014.
Rao RV, Ellerby HM, Bredesen DE. 2004. Coupling endoplasmic reticulum stress to the cell death
program. Cell Death Differ 11: 372–380.
Resendez E Jr, Attenello JW, Grafsky A, Chang CS, Lee AS. 1985. Calcium ionophore A23187
induces expression of glucose-regulated genes and their heterologous fusion genes. Mol Cell Biol 5:
1212–1219.
* Rhind N, Russell P. 2012. Signaling pathways that regulate cell division. Cold Spring Harb Perspect
Biol 4: a005942.
Robinson MJ, Cobb MH. 1997. Mitogen-activated protein kinase pathways. Curr Opin Cell Biol 9:
180–186.
Rodriguez Limardo RG, Ferreiro DN, Roitberg AE, Marti MA, Turjanski AG. 2011. p38γ activation
triggers dynamical changes in allosteric docking sites. Biochemistry 50: 1384–1395.
Ron D, Walter P. 2007. Signal integration in the endoplasmic reticulum unfolded protein response. Nat
Rev Mol Cell Biol 8: 519–529.
Rutkowski DT, Kaufman RJ. 2004. A trip to the ER: Coping with stress. Trends Cell Biol 14: 20–28.
Sabio G, Davis RJ. 2010. cJun NH2-terminal kinase 1 (JNK1): Roles in metabolic regulation of insulin
resistance. Trends Biochem Sci 35: 490–496.
Salvador JM, Mittelstadt PR, Guszczynski T, Copeland TD, Yamaguchi H, Appella E, Fornace AJ Jr,
Ashwell JD. 2005. Alternative p38 activation pathway mediated by T cell receptor-proximal
tyrosine kinases. Nat Immunol 6: 390–395.
* Sassone-Corsi P. 2012. The cyclic AMP pathway. Cold Spring Harb Perspect Biol 4: a011148.
Scull CM, Tabas I. 2011. Mechanisms of ER stress-induced apoptosis in atherosclerosis. Arterioscler
Thromb Vasc Biol 31: 2792–2797.
Seo HY, Kim YD, Lee KM, Min AK, Kim MK, Kim HS, Won KC, Park JY, Lee KU, Choi HS, et al.
2008. Endoplasmic reticulum stress-induced activation of activating transcription factor 6 decreases
insulin gene expression via up-regulation of orphan nuclear receptor small heterodimer partner.
Endocrinology 149: 3832–3841.
Sevier C, Kaiser C. 2008. Ero1 and redox homeostasis in the endoplasmic reticulum. Biochim Biophys
Acta 1783: 549–556.
Sevier C, Cuozzo J, Vala A, Aslund F, Kaiser C. 2001. A flavoprotein oxidase defines a new
endoplasmic reticulum pathway for biosynthetic disulphide bond formation. Nat Cell Biol 3: 874–
882.
Shamu CE, Walter P. 1996. Oligomerization and phosphorylation of the Ire1p kinase during
intracellular signaling from the endoplasmic reticulum to the nucleus. EMBO J 15: 3028–3039.
Sharma NK, Das SK, Mondal AK, Hackney OG, Chu WS, Kern PA, Rasouli N, Spencer HJ, Yao-
Borengasser A, Elbein SC. 2008. Endoplasmic reticulum stress markers are associated with obesity
in nondiabetic subjects. J Clin Endocrinol Metab 93: 4532–4541.
Sharma G, Pallesen J, Das S, Grassucci R, Langlois R, Hampton CM, Kelly DF, des Georges A, Frank
J. 2013. Affinity grid-based cryo-EM of PKC binding to RACK1 on the ribosome. J Struct Biol
181: 190–194.
Shen X, Ellis RE, Lee K, Liu CY, Yang K, Solomon A, Yoshida H, Morimoto R, Kurnit DM, Mori K,
et al. 2001. Complementary signaling pathways regulate the unfolded protein response and are
required for C. elegans development. Cell 107: 893–903.
Shen J, Chen X, Hendershot L, Prywes R. 2002. ER stress regulation of ATF6 localization by
dissociation of BiP/GRP78 binding and unmasking of Golgi localization signals. Dev Cell 3: 99–
111.
Shi Y, Vattem KM, Sood R, An J, Liang J, Stramm L, Wek RC. 1998. Identification and
characterization of pancreatic eukaryotic initiation factor 2 α-subunit kinase, PEK, involved in
translational control. Mol Cell Biol 18: 7499–7509.
Shimazawa M, Ito Y, Inokuchi Y, Hara H. 2007. Involvement of double-stranded RNA-dependent
protein kinase in ER stress-induced retinal neuron damage. Invest Ophthalmol Vis Sci 48: 3729–
3736.
Shiu RP, Pouyssegur J, Pastan I. 1977. Glucose depletion accounts for the induction of two
transformation-sensitive membrane proteins in Rous sarcoma virus-transformed chick embryo
fibroblasts. Proc Natl Acad Sci 74: 3840–3844.
Sidrauski C, Walter P. 1997. The transmembrane kinase Ire1p is a site-specific endonuclease that
initiates mRNA splicing in the unfolded protein response. Cell 90: 1031–1039.
Somers J, Pöyry T, Willis AE. 2013. A perspective on mammalian upstream open reading frame
function. Int J Biochem Cell Biol 45: 1690–1700.
Sumara G, Formentini I, Collins S, Sumara I, Windak R, Bodenmiller B, Ramracheya R, Caille D,
Jiang H, Platt KA, et al. 2009. Regulation of PKD by the MAPK p38δ in insulin secretion and
glucose homeostasis. Cell 136: 235–248.
Szegezdi E, Logue SE, Gorman AM, Samali A. 2006. Mediators of endoplasmic reticulum stress-
induced apoptosis. EMBO Rep 7: 880–885.
Takekawa M, Saito H. 1998. A family of stress-inducible GADD45-like proteins mediate activation of
the stress-responsive MTK1/MEKK4 MAPKKK. Cell 95: 521–530.
Talukdar S, Olefsky JM, Osborn O. 2011. Targeting GPR120 and other fatty acid-sensing GPCRs
ameliorates insulin resistance and inflammatory diseases. Trends Pharm Sci 32: 543–550.
Tanoue T, Nishida E. 2003. Molecular recognitions in the MAP kinase cascades. Cell Signal 15: 455–
462.
Tournier C, Hess P, Yang DD, Xu J, Turner TK, Nimnual A, Bar-Sagi D, Jones SN, Flavell RA, Davis
RJ. 2000. Requirement of JNK for stress-induced activation of the cytochrome c-mediated death
pathway. Science 288: 870–874.
Tournier C, Dong C, Turner TK, Jones SN, Flavell RA, Davis RJ. 2001. MKK7 is an essential
component of the JNK signal transduction pathway activated by proinflammatory cytokines. Genes
Dev 15: 1419–1426.
Trusina A, Papa FR, Tang C. 2008. Rationalizing translation attenuation in the network architecture of
the unfolded protein response. Proc Natl Acad Sci 105: 20280–20285.
Tu BP, Weissman JS. 2002. The FAD- and O2-dependent reaction cycle of Ero1-mediated oxidative
protein folding in the endoplasmic reticulum. Mol Cell 10: 983–994.
Uehara T, Nakamura T, Yao D, Shi ZQ, Gu Z, Ma Y, Masliah E, Nomura Y, Lipton SA. 2006. S-
nitrosylated protein-disulphide isomerase links protein misfolding to neurodegeneration. Nature
441: 513–517.
Urano F, Wang X, Bertolotti A, Zhang Y, Chung P, Harding HP, Ron D. 2000. Coupling of stress in
the ER to activation of JNK protein kinases by transmembrane protein kinase IRE1. Science 287:
664–666.
Ventura JJ, Cogswell P, Flavell RA, Baldwin AS Jr, Davis RJ. 2004. JNK potentiates TNF-stimulated
necrosis by increasing the production of cytotoxic reactive oxygen species. Genes Dev 18: 2905–
2915.
Ventura JJ, Hubner A, Zhang C, Flavell RA, Shokat KM, Davis RJ. 2006. Chemical genetic analysis of
the time course of signal transduction by JNK. Mol Cell 21: 701–710.
Walter P, Ron D. 2011. The unfolded protein response: From stress pathway to homeostatic regulation.
Science 334: 1081–1086.
Wang XZ, Harding HP, Zhang Y, Jolicoeur EM, Kuroda M, Ron D. 1998. Cloning of mammalian Ire1
reveals diversity in the ER stress responses. EMBO J 17: 5708–5717.
* Ward PS, Thompson CB. 2012. Signaling in control of cell growth and metabolism. Cold Spring
Harb Perspect Biol 4: a006783.
Wei Y, Pattingre S, Sinha S, Bassik M, Levine B. 2008. JNK1-mediated phosphorylation of Bcl-2
regulates starvation-induced autophagy. Mol Cell 30: 678–688.
Welch WJ, Garrels JI, Thomas GP, Lin JJ, Feramisco JR. 1983. Biochemical characterization of the
mammalian stress proteins and identification of two stress proteins as glucose- and Ca2+-
ionophore-regulated proteins. J Biol Chem 258: 7102–7111.
Whinston E, Omerza G, Singh A, Tio CW, Winter E. 2013. Activation of the Smk1 mitogen-activated
protein kinase by developmentally regulated autophosphorylation. Mol Cell Biol 33: 688–700.
Whitmarsh AJ, Davis RJ. 2000. Regulation of transcription factor function by phosphorylation. Cell
Mol Life Sci 57: 1172–1183.
Whitmarsh AJ, Kuan CY, Kennedy NJ, Kelkar N, Haydar TF, Mordes JP, Appel M, Rossini AA, Jones
SN, Flavell RA, et al. 2001. Requirement of the JIP1 scaffold protein for stress-induced JNK
activation. Genes Dev 15: 2421–2432.
Wiggins CM, Tsvetkov P, Johnson M, Joyce CL, Lamb CA, Bryant NJ, Komander D, Shaul Y, Cook
SJ. 2011. BIMEL, an intrinsically disordered protein, is degraded by 20S proteasomes in the
absence of poly-ubiquitylation. J Cell Sci 124: 969–977.
Witzel F, Maddison L, Blüthgen N. 2012. How scaffolds shape MAPK signaling: What we know and
opportunities for systems approaches. Front Physiol 3: 475.
Wong HK, Fricker M, Wyttenbach A, Villunger A, Michalak EM, Strasser A, Tolkovsky AM. 2005.
Mutually exclusive subsets of BH3-only proteins are activated by the p53 and c-Jun N-terminal
kinase/c-Jun signaling pathways during cortical neuron apoptosis induced by arsenite. Mol Cell Biol
25: 8732–8747.
Wu N, Zheng B, Shaywitz A, Dagon Y, Tower C, Bellinger G, Shen CH, Wen J, Asara J, McGraw TE,
et al. 2013 AMPK-dependent degradation of TXNIP upon energy stress leads to enhanced glucose
uptake via GLUT1. Mol Cell 49: 1167–1175.
Xia Z, Dickens M, Raingeaud J, Davis RJ, Greenberg ME. 1995. Opposing effects of ERK and JNK-
p38 MAP kinases on apoptosis. Science 270: 1326–1331.
Xiao C, Giacca A, Lewis GF. 2011. Sodium phenylbutyrate, a drug with known capacity to reduce
endoplasmic reticulum stress, partially alleviates lipid-induced insulin resistance and β-cell
dysfunction in humans. Diabetes 60: 918–924.
Yamada T, Ishihara H, Tamura A, Takahashi R, Yamaguchi S, Takei D, Tokita A, Satake C, Tashiro F,
Katagiri H, et al. 2006. WFS1-deficiency increases endoplasmic reticulum stress, impairs cell cycle
progression and triggers the apoptotic pathway specifically in pancreatic β-cells. Hum Mol Genet
15: 1600–1609.
Yamazaki H, Hiramatsu N, Hayakawa K, Tagawa Y, Okamura M, Ogata R, Huang T, Yao J, Paton A,
Paton J, et al. 2009. Activation of the Akt-NF-κB pathway by subtilase cytotoxin through the ATF6
branch of the unfolded protein response. J Immunol 183: 1480–1487.
Yang DD, Kuan CY, Whitmarsh AJ, Rincon M, Zheng TS, Davis RJ, Rakic P, Flavell RA. 1997.
Absence of excitotoxicity-induced apoptosis in the hippocampus of mice lacking the Jnk3 gene.
Nature 389: 865–870.
Yecies JL, Manning BD. 2011. mTOR links oncogenic signaling to tumor cell metabolism. J Mol Med
89: 221–228.
Yoneda T, Imaizumi K, Oono K, Yui D, Gomi F, Katayama T, Tohyama M. 2001. Activation of
caspase-12, an endoplastic reticulum (ER) resident caspase, through tumor necrosis factor receptor-
associated factor 2-dependent mechanism in response to the ER stress. J Biol Chem 276: 13935–
13940.
Yoshida H, Haze K, Yanagi H, Yura T, Mori K. 1998. Identification of the cis-acting endoplasmic
reticulum stress response element responsible for transcriptional induction of mammalian glucose-
regulated proteins. Involvement of basic leucine zipper transcription factors. J Biol Chem 273:
33741–33749.
Yoshida H, Matsui T, Yamamoto A, Okada T, Mori K. 2001. XBP1 mRNA is induced by ATF6 and
spliced by IRE1 in response to ER stress to produce a highly active transcription factor. Cell 107:
881–891.
Yoshida H, Matsui T, Hosokawa N, Kaufman RJ, Nagata K, Mori K. 2003. A time-dependent phase
shift in the mammalian unfolded protein response. Dev Cell 4: 265–271.
Zhang P, McGrath B, Li S, Frank A, Zambito F, Reinert J, Gannon M, Ma K, McNaughton K, Cavener
DR. 2002. The PERK eukaryotic initiation factor 2 α kinase is required for the development of the
skeletal system, postnatal growth, and the function and viability of the pancreas. Mol Cell Biol 22:
3864–3874.
Zhang X, Zhang G, Zhang H, Karin M, Bai H, Cai D. 2008. Hypothalamic IKKβ/NF-κB and ER stress
link overnutrition to energy imbalance and obesity. Cell 135: 61–73.
Zhao L, Ackerman SL. 2006. Endoplasmic reticulum stress in health and disease. Curr Opin Cell Biol
18: 444–452.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a006072
CHAPTER 19
SUMMARY
Outline
1 Introduction
2 Type I cell death: Apoptosis
3 Type II cell death and autophagy
4 Type III cell death: Necrosis
5 Beyond type III—Other forms of cell death
6 Conclusion
References
1 INTRODUCTION
Although cell death can happen as a result of overwhelming damage, most
cell deaths in animals occur in an active manner, as a consequence of specific
signaling events. In general, there are three types of cell death, defined in
large part by the appearance of the dying cell: apoptosis (also known as type I
cell death), autophagic cell death (type II), and necrosis (type III) (Galluzzi et
al. 2007).
Apoptosis is characterized by cell shrinkage, membrane blebbing, and
condensation of the chromatin (pyknosis) (Kerr et al. 1972). It can be further
defined as cell death accompanied by the activation of caspase proteases
(Galluzzi et al. 2012). Two major signaling pathways trigger apoptotic cell
death: the mitochondrial (the intrinsic) pathway and the death receptor (the
extrinsic) pathway. The latter involves a classical ligand–cell-surface-
receptor interaction. For example, cytotoxic lymphocytes can kill infected or
transformed cells by expressing ligands for death receptors (DRs), a subset of
the tumor necrosis factor (TNF) receptor (TNFR) family. These ligands
induce apoptotic cell death of the targeted cells provided they express such
DRs. DR-induced cell death in general is critical for immune system function
and homeostasis. In contrast, the mitochondrial apoptotic pathway is usually
initiated in a cell-autonomous manner. Most cellular stresses, such as DNA
damage (induced by genotoxic agents or defects in DNA repair) or
endoplasmic reticulum (ER) stress (induced by the accumulation of unfolded
proteins), actively engage apoptosis when cells are damaged beyond repair.
Conversely, the lack of a signal, such as those activated by growth factors
(e.g., cytokines and neurotrophic factors), can lead to cell death. This
mechanism is critical for the development of the nervous system in
vertebrates and it is estimated that half of the neurons generated die during
this process (Buss et al. 2006). This cell death is due, in part, to the failure of
some neuronal precursors to properly migrate or innervate their targets and
the consequent lack of neurotrophic factor stimulation. Similarly, during an
immune response, cytokine deprivation (together with the DR pathway) is
responsible for the acute contraction of the lymphocyte population after
clearance of the pathogen. Another example of “loss-of-signal”-induced cell
death is the particular form of apoptosis called anoikis, which occurs when
epithelial or endothelial cells detach from the extracellular matrix (ECM). In
this scenario, unligated ECM receptors of the integrin family cease to induce
prosurvival signaling pathways, eventually leading to apoptosis. This
mechanism prevents cells shedding from their original location from
colonizing elsewhere (a characteristic of metastatic cancer cells). Finally,
apoptosis can be induced by oncogenes (e.g., Myc) as a safeguard mechanism
against cancer development. This process is controlled in part by a p53-
dependent apoptotic pathway, which is activated in response to aberrant
mitogenic signals resulting from oncogene overexpression or mutation. As a
consequence, evasion of apoptotic cell death is often a requisite to sustain
oncogene transformation (see Ch. 21 [Sever and Brugge 2014]).
Autophagic cell death is characterized by the appearance of large
intracellular vesicles and engagement of the autophagy machinery. Note that
although autophagy (i.e., the membrane engulfment and catabolic
degradation of parts of the cytoplasm) is a well-defined process, its function
as an active cell death mechanism remains highly controversial. Autophagy is
mainly a survival process engaged in response to a metabolic crisis (e.g., low
ATP levels and nutrient and amino acid deprivation) or to remove damaged
organelles (e.g., mitochondria with low membrane potential) and protein
aggregates. As a stress response, autophagy accompanies rather than
promotes cell death in most scenarios and merely represents a failed survival
attempt (Shen et al. 2012). Nevertheless, there are specific examples in which
the autophagy machinery is absolutely required for cell death. During
Drosophila metamorphosis, obsolete larval tissues such as the midgut and
salivary glands regress through massive autophagic cell death, a process
triggered by the steroid hormone ecdysone. In this particular case, a
deficiency in genes of the autophagic signaling pathway alters the cell death
program (Berry and Baehrecke 2007; Denton et al. 2009). Autophagic cell
death has also been reported in response to deregulated H-Ras activity and
could therefore represent a safeguard mechanism against oncogenic
transformation (Elgendy et al. 2011).
Necrosis is characterized by cell swelling and plasma membrane rupture,
and a loss of organellar structure without chromatin condensation. Although
necrosis can occur as a consequence of irreparable cell damage, at least one
pathway of active necrosis exists. This form of cell death, sometimes called
necroptosis, is engaged by several signaling pathways that all converge on
the activation of receptor-interacting protein kinase 3 (RIP3). RIP3 is
activated upon recruitment to macromolecular complexes downstream from
various cell-surface receptors: DRs, Toll-like receptors (TLRs), and the T-
cell receptor (TCR). Additionally, DNA damage can directly induce the
formation of a RIP3-activation platform, independently of cell-surface
receptor ligation. Finally, RIP3-dependent necrosis is also triggered by the
cytosolic DNA sensor, DNA-dependent activator of interferon (DAI)
regulatory factors, following virus infection and the presence, in the cytosol,
of double-stranded viral DNA.
Here we are concerned with signaling leading to cell death in vertebrates,
and focus on processes that are at least partially understood at the molecular
level. We do not discuss the physiology and pathology of cell death in detail
or processes involved in the clearance of dying cells. Readers will find a
more complete overview of these topics and other aspects of cell death
elsewhere (Green 2011b).
Figure 1. The caspase protein family. Initiator caspases (caspase-2, caspase-8, and caspase-9) are the
apical caspases of the apoptotic-signaling cascade. Initiator caspases are produced as inactive zymogens
composed of a prodomain (containing a CARD or a death effector domain [DED]) and a large and
small subunit. They are recruited through their prodomains into large activation platforms and activated
by dimerization. In contrast, executioner caspases (caspase-3, caspase-6, and caspase-7) are activated
by cleavage of the zymogen between the large and small subunits and are therefore dependent on
initiator caspases for their activation. Catalytically active caspases are composed of a heterotetramer of
two small and two large subunits.
Figure 3. Regulation of mitochondrial outer membrane integrity by the Bcl2 protein family. (A)
Members of the Bcl2 protein family are characterized by the presence of one or more Bcl2 homology
(BH) region. The antiapoptotic Bcl2 proteins (e.g., Bcl2, Bcl-xL, and Mcl1) and the proapoptotic
effectors (e.g., Bax and Bak) share four BH regions and a similar globular structure. BH3-only proteins
(e.g., Bid, Bim, Bad, and Noxa) are characterized by a single BH region (BH3). (B) The proapoptotic
effectors reside in cells in inactive forms tethered to the outer mitochondrial membrane (Bak) or soluble
in the cytosol (Bax). Upon activation, Bax and Bak oligomerize and further insert into the membrane,
causing mitochondrial outer membrane permeabilization (MOMP) and thus apoptosis. (C) Bax/Bak
activation is triggered following the transient binding of a subset of direct activator BH3-only proteins
(e.g., Bid and Bim). Antiapoptotic Bcl2 proteins inhibit MOMP by sequestering the direct activator
proteins and/or the effectors. Another group of BH3-only proteins, called sensitizers or derepressors
(e.g., Bad and Noxa) promote MOMP by antagonizing antiapoptotic Bcl2 proteins, thereby releasing
both direct activator BH3-only proteins and Bax/Bak.
Bax and Bak are directly responsible for the loss of mitochondrial outer
membrane integrity (Fig. 3B). Upon activation, they form large oligomers
that insert into the mitochondrial outer membrane, disrupting it (Eskes et al.
2000; Korsmeyer et al. 2000; Dewson et al. 2008, 2009). The precise nature
of the disruption remains unclear, but it allows the near simultaneous release
of all intermembrane space proteins (Goldstein et al. 2000; Munoz-Pinedo et
al. 2006).
Bax and Bak act redundantly in MOMP and at least one of them is
required to permeabilize mitochondria. In living cells these proteins are
generally inactive, but become activated in response to upstream events. At
least two of the BH3-only proteins (Bim and active Bid) activate Bax and
Bak through transient interaction (Fig. 3C), although other conditions (such
as heat, changes in pH, and changes in the lipid milieu) may activate the
effectors independently of BH3-only proteins (Wei et al. 2000; Kuwana et al.
2002, 2005; Letai et al. 2002).
MOMP is antagonized by the antiapoptotic Bcl2 proteins, which bind to
and inhibit both Bax/Bak and the BH3-only proteins by interacting with their
BH3 domains (Llambi et al. 2011) (Fig. 3C). Antiapoptotic Bcl2 proteins are
also regulated at both the transcriptional and posttranslational levels. In
particular, Mcl1 degradation by the ubiquitin-proteasome system participates
in apoptosis induction following several cellular stresses. Upon DNA
damage, the BH3-domain-containing protein Mcl1 ubiquitin ligase E3
(MULE) directly binds to Mcl1 and catalyzes its ubiquitylation and
degradation (Warr et al. 2005). During antitubulin chemotherapeutic-induced
mitotic arrest, Mcl1 is phosphorylated by stress-activated and mitotic kinases
such as the MAP kinase (MAPK) Jun amino-terminal kinase (JNK), casein
kinase II (CKII), and p38 MAPK. These phosphorylation events unveil a
degron on Mcl1 that recruits the SCF-Fbw7 ubiquitin ligase complex, thus
targeting Mcl1 for ubiquitylation and degradation (Wertz et al. 2011).
Similarly, Mcl1 is degraded following growth factor withdrawal (e.g., IL3)
and the subsequent loss of phosphoinositide 3-kinase (PI3K)-Akt signaling.
This process relieves glycogen synthase kinase 3 (GSK3) from Akt
inhibition. GSK3 then phosphorylates Mcl1, allowing its ubiquitylation by
the E3 ligase β-transductin-repeat-containing protein (β-TrCP) and its
subsequent degradation (Maurer et al. 2006; Ding et al. 2007). Conversely,
cancer cells can increase Mcl1 stability and overall resistance to cellular
stress by expressing USP9X, a deubiquitylase that removes polyubiquitin
chains from Mcl1 (Schwickart et al. 2010).
Proteins of the Bcl2 family integrate pro- and antiapoptotic signals in
healthy and stressed cells and therefore constitute one of the main signaling
nodes in the life or death decision. Within this family, BH3-only proteins
constitute the main upstream sensors of the mitochondrial apoptotic pathway.
A wide variety of signaling pathways converge on the BH3-only family of
proteins and regulate their expression level and activity both transcriptionally
and posttranscriptionally (see Table 1). For example, Bid is activated upon
cleavage by caspase-8 following DR ligation (Li et al. 1998, Luo et al. 1998).
In doing so, Bid coordinates the cross-regulation between the extrinsic and
intrinsic apoptotic pathways. Additionally, proteolytic activation of Bid can
be achieved by granzyme B during the cytotoxic lymphocyte killing process
(Heibein et al. 2000; Sutton et al. 2000). The apoptotic response to genotoxic
stress is performed, in part, by Puma and Noxa (Jeffers et al. 2003; Villunger
et al. 2003). Both are direct transcriptional targets of the tumor suppressor
p53 (Oda et al. 2000; Nakano and Vousden 2001; Yu et al. 2003), although
other stimuli regulate their expression levels (see Table 1). Similarly, Bim is
transcriptionally up-regulated by the forkhead transcription factor FOXO3A
upon cytokine deprivation (Dijkers et al. 2000; Gilley et al. 2003; Urbich et
al. 2005). Bim is also a major apoptotic factor of the ER stress pathway
engaged in response to accumulation of unfolded proteins (Puthalakath et al.
2007). Finally, Bad activity is negatively regulated through phosphorylation
by several kinases, such as Akt (Datta et al. 1997; del Peso et al. 1997),
which induces its sequestration by 14-3-3 proteins. Upon growth factor
deprivation and loss of Akt signaling, Bad is released and antagonizes
antiapoptotic Bcl2 proteins (Zha et al. 1996; Datta et al. 1997).
6 CONCLUSION
Active or programmed cell death is essential to maintain homeostasis in
multicellular organisms as well as for the selective elimination of potentially
harmful or infected cells. Accordingly, deregulation of the signaling
pathways that trigger cell death can lead to the development of catastrophic
diseases such as cancer and autoimmunity (too little cell death) as well as
degenerative diseases (too much cell death). Therefore, the existence of
tightly controlled and efficient means to induce cell death can be interpreted
as the logical consequence of the evolution of multicellular organisms.
However, the necessity for so many different death-signaling pathways might
appear counterintuitive. Taken together, cell death induction could be viewed
as a simple signaling process with multiple inputs/stimuli and one outcome:
the death of the cell. However, the way by which a cell dies has important
consequences for neighboring cells and sometimes the entire organism. For
example, apoptotic and necrotic cells display divergent inflammatory
properties and trigger different immune responses. Additionally, particular
death programs include the release of proliferative signals that trigger
compensatory proliferation in surrounding tissues. These signals might differ
from one type of cell death to the next. Finally, death-signaling pathways are
clearly interconnected. For example, autophagic cell death is often
potentiated by caspase activation, whereas RIP-dependent necrosis is
antagonized by a caspase-dependent activity. Cross talk between these
pathways potentially provides numerous backup mechanisms for cell death
programs and could explain why inhibition of a single program often has
minor consequences for the organism. A better understanding of the impact
of each type of cell death on surrounding tissues and of the interplay between
these cell death programs might help to answer some of these questions.
REFERENCES
*Reference is in this book.
Akiyama T, Bouillet P, Miyazaki T, Kadono Y, Chikuda H, Chung U-I, Fukuda A, Hikita A, Seto H,
Okada T, et al. 2003. Regulation of osteoclast apoptosis by ubiquitylation of proapoptotic BH3-only
Bcl-2 family member Bim. EMBO J 22: 6653–6664.
Allan LA, Clarke PR. 2007. Phosphorylation of caspase-9 by CDK1/cyclin B1 protects mitotic cells
against apoptosis. Mol Cell 26: 301–310.
Allan LA, Clarke PR. 2009. Apoptosis and autophagy: Regulation of caspase 9 by phosphorylation.
FEBS J 276: 6063–6073.
Allan LA, Morrice N, Brady S, Magee G, Pathak S, Clarke PR. 2003. Inhibition of caspase-9 through
phosphorylation at Thr 125 by ERK MAPK. Nat Cell Biol 5: 647–654.
Alves NL, Derks IA, Berk E, Spijker R, van Lier RA, Eldering E. 2006. The Noxa/Mcl-1 axis regulates
susceptibility to apoptosis under glucose limitation in dividing T cells. Immunity 24: 703–716.
Andersen JL, Johnson CE, Freel CD, Parrish AB, Day JL, Buchakjian MR, Nutt LK, Thompson JW,
Moseley MA, Kornbluth S. 2009. Restraint of apoptosis during mitosis through interdomain
phosphorylation of caspase 2. EMBO J 28: 3216–3227.
Andersen JL, Thompson JW, Lindblom KR, Johnson ES, Yang CS, Lilley LR, Freel CD, Moseley MA,
Kornbluth S. 2011. A biotin switch-based proteomics approach identifies 14–3-3ζ as a target of
Sirt1 in the metabolic regulation of caspase 2. Mol Cell 43: 834–842.
Ayllón V, Martínez-A C, García A, Cayla X, Rebollo A. 2000. Protein phosphatase 1α is a Ras-
activated Bad phosphatase that regulates interleukin-2 deprivation-induced apoptosis. EMBO J 19:
2237–2246.
Baines CP, Kaiser RA, Purcell NH, Blair NS, Osinska H, Hambleton MA, Brunskill EW, Sayen MR,
Gottlieb RA, Dorn GW, et al. 2005. Loss of cyclophilin D reveals a critical role for mitochondrial
permeability transition in cell death. Nature 434: 658–662.
Baliga BC, Read SH, Kumar S. 2004. The biochemical mechanism of caspase 2 activation. Cell Death
Differ 11: 1234–1241.
Basso E, Fante L, Fowlkes J, Petronilli V, Forte MA, Bernardi P. 2005. Properties of the permeability
transition pore in mitochondria devoid of Cyclophilin D. J Biol Chem 280: 18558–18561.
Bellot G, Garcia-Medina R, Gounon P, Chiche J, Roux D, Pouyssegur J, Mazure NM. 2009. Hypoxia-
induced autophagy is mediated through hypoxia-inducible factor induction of BNIP3 and BNIP3L
via their BH3 domains. Mol Cell Biol 29: 2570–2581.
Bender CE, Fitzgerald P, Tait SWG, Llambi F, McStay GP, Tupper DO, Pellettieri J, Sánchez Alvarado
A, Salvesen GS, Green DR. 2012. Mitochondrial pathway of apoptosis is ancestral in metazoans.
Proc Natl Acad Sci 109: 4904–4909.
Bergsbaken T, Cookson BT. 2007. Macrophage activation redirects yersinia-infected host cell death
from apoptosis to caspase-1-dependent pyroptosis. PLoS Pathog 3: e161.
Berry DL, Baehrecke EH. 2007. Growth arrest and autophagy are required for salivary gland cell
degradation in Drosophila. Cell 131: 1137–1148.
Bonzon C, Bouchier-Hayes L, Pagliari LJ, Green DR, Newmeyer DD. 2006. Caspase-2-induced
apoptosis requires bid cleavage: A physiological role for bid in heat shock-induced death. Mol Biol
Cell 17: 2150–2157.
Bouchier-Hayes L, Oberst A, McStay GP, Connell S, Tait SWG, Dillon CP, Flanagan JM, Beere HM,
Green DR. 2009. Characterization of cytoplasmic caspase-2 activation by induced proximity. Mol
Cell 35: 830–840.
Bratton SB, Salvesen GS. 2010. Regulation of the Apaf-1-caspase-9 apoptosome. J Cell Sci 123: 3209–
3214.
Bratton SB, Walker G, Srinivasula SM, Sun XM, Butterworth M, Alnemri ES, Cohen GM. 2001.
Recruitment, activation and retention of caspases 9 and 3 by Apaf-1 apoptosome and associated
XIAP complexes. EMBO J 20: 998–1009.
Brennan MA, Cookson BT. 2000. Salmonella induces macrophage death by caspase-1-dependent
necrosis. Mol Microbiol 38: 31–40.
Brinkmann K, Zigrino P, Witt A, Schell M, Ackermann L, Broxtermann P, Schull S, Andree M,
Coutelle O, Yazdanpanah B, et al. 2013. Ubiquitin C-terminal hydrolase-L1 potentiates cancer
chemosensitivity by stabilizing NOXA. Cell Rep 3: 881–891.
Bruick RK. 2000. Expression of the gene encoding the proapoptotic Nip3 protein is induced by
hypoxia. Proc Natl Acad Sci 97: 9082–9087.
Buss RR, Sun W, Oppenheim RW. 2006. Adaptive roles of programmed cell death during nervous
system development. Annu Rev Neurosci 29: 1–35.
Cai Z, Jitkaew S, Zhao J, Chiang HC, Choksi S, Liu J, Ward Y, Wu LG, Liu ZG. 2014. Plasma
membrane translocation of trimerized MLKL protein is required for TNF-induced necroptosis. Nat
Cell Biol 16: 55–65.
Castedo M, Perfettini JL, Roumier T, Andreau K, Medema R, Kroemer G. 2004. Cell death by mitotic
catastrophe: A molecular definition. Oncogene 23: 2825–2837.
Chen M, He H, Zhan S, Krajewski S, Reed JC, Gottlieb RA. 2001. Bid is cleaved by calpain to an
active fragment in vitro and during myocardial ischemia/reperfusion. J Biol Chem 276: 30724–
30728.
Ch’en IL, Tsau JS, Molkentin JD, Komatsu M, Hedrick SM. 2011. Mechanisms of necroptosis in T
cells. J Exp Med 208: 633–641.
Chen X, Li W, Ren J, Huang D, He WT, Song Y, Yang C, Zheng X, Chen P, Han J. 2014.
Translocation of mixed lineage kinase domain-like protein to plasma membrane leads to necrotic
cell death. Cell Res 24: 105–121.
Chiang CW, Harris G, Ellig C, Masters SC, Subramanian R, Shenolikar S, Wadzinski BE, Yang E.
2001. Protein phosphatase 2A activates the proapoptotic function of BAD in interleukin- 3-
dependent lymphoid cells by a mechanism requiring 14–3-3 dissociation. Blood 97: 1289–1297.
Chinnaiyan AM, Chaudhary D, O’Rourke K, Koonin EV, Dixit VM. 1997a. Role of CED-4 in the
activation of CED-3. Nature 388: 728–729.
Chinnaiyan AM, O’Rourke K, Lane BR, Dixit VM. 1997b. Interaction of CED-4 with CED-3 and
CED-9: A molecular framework for cell death. Science 275: 1122–1126.
Chipuk JE, Moldoveanu T, Llambi F, Parsons MJ, Green DR. 2010. The BCL-2 family reunion. Mol
Cell 37: 299–310.
Cho YS, Challa S, Moquin D, Genga R, Ray TD, Guildford M, Chan FK. 2009. Phosphorylation-
driven assembly of the RIP1-RIP3 complex regulates programmed necrosis and virus-induced
inflammation. Cell 137: 1112–1123.
Chou JJ, Matsuo H, Duan H, Wagner G. 1998. Solution structure of the RAIDD CARD and model for
CARD/CARD interaction in caspase-2 and caspase-9 recruitment. Cell 94: 171–180.
Christofferson DE, Yuan J. 2010. Necroptosis as an alternative form of programmed cell death. Curr
Opin Cell Biol 22: 263–268.
Coleman ML, Sahai EA, Yeo M, Bosch M, Dewar A, Olson MF. 2001. Membrane blebbing during
apoptosis results from caspase-mediated activation of ROCK I. Nature Cell Biol 3: 339–345.
Conradt B, Horvitz HR. 1998. The C. elegans protein EGL-1 is required for programmed cell death and
interacts with the Bcl-2-like protein CED-9. Cell 93: 519–529.
Cookson BT, Brennan MA. 2001. Proinflammatory programmed cell death. Trends Microbiol 9: 113–
114.
Cotteret S, Jaffer ZM, Beeser A, Chernoff J. 2003. p21-Activated kinase 5 (Pak5) localizes to
mitochondria and inhibits apoptosis by phosphorylating BAD. Mol Cell Biol 23: 5526–5539.
Crawford ED, Wells JA. 2011. Caspase substrates and cellular remodeling. Annu Rev Biochem 80:
1055–1087.
Datta SR, Dudek H, Tao X, Masters S, Fu H, Gotoh Y, Greenberg ME. 1997. Akt phosphorylation of
BAD couples survival signals to the cell-intrinsic death machinery. Cell 91: 231–241.
Datta SR, Katsov A, Hu L, Petros A, Fesik SW, Yaffe MB, Greenberg ME. 2000. 14-3-3 proteins and
survival kinases cooperate to inactivate BAD by BH3 domain phosphorylation. Mol Cell 6: 41–51.
Datta SR, Ranger AM, Lin MZ, Sturgill JF, Ma Y-C, Cowan CW, Dikkes P, Korsmeyer SJ, Greenberg
ME. 2002. Survival factor-mediated BAD phosphorylation raises the mitochondrial threshold for
apoptosis. Dev Cell 3: 631–643.
Declercq W, Vanden Berghe T, Vandenabeele P. 2009. RIP kinases at the crossroads of cell death and
survival. Cell 138: 229–232.
Degterev A, Huang Z, Boyce M, Li Y, Jagtap P, Mizushima N, Cuny GD, Mitchison TJ, Moskowitz
MA, Yuan J. 2005. Chemical inhibitor of nonapoptotic cell death with therapeutic potential for
ischemic brain injury. Nat Chem Biol 1: 112–119.
Degterev A, Hitomi J, Germscheid M, Ch’en IL, Korkina O, Teng X, Abbott D, Cuny GD, Yuan C,
Wagner G, et al. 2008. Identification of RIP1 kinase as a specific cellular target of necrostatins. Nat
Chem Biol 4: 313–321.
Dehan E, Bassermann F, Guardavaccaro D, Vasiliver-Shamis G, Cohen M, Lowes KN, Dustin M,
Huang DCS, Taunton J, Pagano M. 2009. βTrCP- and Rsk1/2-mediated degradation of BimEL
inhibits apoptosis. Mol Cell 33: 109–116.
del Peso L, González-García M, Page C, Herrera R, Nuñez G. 1997. Interleukin-3-induced
phosphorylation of BAD through the protein kinase Akt. Science 278: 687–689.
Denton D, Shravage B, Simin R, Mills K, Berry DL, Baehrecke EH, Kumar S. 2009. Autophagy, not
apoptosis, is essential for midgut cell death in Drosophila. Curr Biol 19: 1741–1746.
* Devreotes P, Horwitz AR. 2014. Signaling networks that regulate cell migration. Cold Spring Harb
Perspect Biol doi: 10.1101/cshperspect.a005959.
Dewson G, Kratina T, Sim HW, Puthalakath H, Adams JM, Colman PM, Kluck RM. 2008. To trigger
apoptosis, Bak exposes its BH3 domain and homodimerizes via BH3:Groove interactions. Mol Cell
30: 369–380.
Dewson G, Kratina T, Czabotar P, Day CL, Adams JM, Kluck RM. 2009. Bak activation for apoptosis
involves oligomerization of dimers via their α6 helices. Mol Cell 36: 696–703.
Di Bartolomeo S, Corazzari M, Nazio F, Oliverio S, Lisi G, Antonioli M, Pagliarini V, Matteoni S,
Fuoco C, Giunta L, et al. 2010. The dynamic interaction of AMBRA1 with the dynein motor
complex regulates mammalian autophagy. J Cell Biol 191: 155–168.
Dickens LS, Powley IR, Hughes MA, MacFarlane M. 2012. The “complexities” of life and death:
Death receptor signalling platforms. Exp Cell Res 318: 1269–1277.
Dijkers PF, Medema RH, Lammers JW, Koenderman L, Coffer PJ. 2000. Expression of the
proapoptotic Bcl-2 family member Bim is regulated by the forkhead transcription factor FKHR-L1.
Curr Biol 10: 1201–1204.
Ding Q, He X, Hsu JM, Xia W, Chen CT, Li LY, Lee DF, Liu JC, Zhong Q, Wang X, et al. 2007.
Degradation of Mcl-1 by β-TrCP mediates glycogen synthase kinase 3-induced tumor suppression
and chemosensitization. Mol Cell Biol 27: 4006–4017.
Dixon SJ, Lemberg KM, Lamprecht MR, Skouta R, Zaitsev EM, Gleason CE, Patel DN, Bauer AJ,
Cantley AM, Yang WS, et al. 2012. Ferroptosis: An iron-dependent form of nonapoptotic cell
death. Cell 149: 1060–1072.
Donovan N, Becker EBE, Konishi Y, Bonni A. 2002. JNK phosphorylation and activation of BAD
couples the stress-activated signaling pathway to the cell death machinery. J Biol Chem 277:
40944–40949.
Duan H, Dixit VM. 1997. RAIDD is a new “death” adaptor molecule. Nature 385: 86–89.
Eckelman BP, Salvesen GS, Scott FL. 2006. Human inhibitor of apoptosis proteins: Why XIAP is the
black sheep of the family. EMBO Rep 7: 988–994.
Egan DF, Shackelford DB, Mihaylova MM, Gelino S, Kohnz RA, Mair W, Vasquez DS, Joshi A,
Gwinn DM, Taylor R, et al. 2011. Phosphorylation of ULK1 (hATG1) by AMP-activated protein
kinase connects energy sensing to mitophagy. Science 331: 456–461.
Ekoff M, Kaufmann T, Engstrom M, Motoyama N, Villunger A, Jonsson JI, Strasser A, Nilsson G.
2007. The BH3-only protein Puma plays an essential role in cytokine deprivation induced apoptosis
of mast cells. Blood 110: 3209–3217.
Elgendy M, Sheridan C, Brumatti G, Martin SJ. 2011. Oncogenic Ras-induced expression of Noxa and
Beclin-1 promotes autophagic cell death and limits clonogenic survival. Mol Cell 42: 23–35.
Enari M, Sakahira H, Yokoyama H, Okawa K, Iwamatsu A, Nagata S. 1998. A caspase-activated
DNase that degrades DNA during apoptosis, and its inhibitor ICAD. Nature 391: 43–50.
Erlacher M, Michalak EM, Kelly PN, Labi V, Niederegger H, Coultas L, Adams JM, Strasser A,
Villunger A. 2005. BH3-only proteins Puma and Bim are rate-limiting for γ-radiation- and
glucocorticoid-induced apoptosis of lymphoid cells in vivo. Blood 106: 4131–4138.
Eskelinen EL, Illert AL, Tanaka Y, Schwarzmann G, Blanz J, Von Figura K, Saftig P. 2002. Role of
LAMP-2 in lysosome biogenesis and autophagy. Mol Biol Cell 13: 3355–3368.
Eskelinen EL, Schmidt CK, Neu S, Willenborg M, Fuertes G, Salvador N, Tanaka Y, Lullmann-Rauch
R, Hartmann D, Heeren J, et al. 2004. Disturbed cholesterol traffic but normal proteolytic function
in LAMP-1/LAMP-2 double-deficient fibroblasts. Mol Biol Cell 15: 3132–3145.
Eskes R, Desagher S, Antonsson B, Martinou JC. 2000. Bid induces the oligomerization and insertion
of Bax into the outer mitochondrial membrane. Mol Cell Biol 20: 929–935.
Faustin B, Lartigue L, Bruey JM, Luciano F, Sergienko E, Bailly-Maitre B, Volkmann N, Hanein D,
Rouiller I, Reed JC. 2007. Reconstituted NALP1 inflammasome reveals two-step mechanism of
caspase-1 activation. Mol Cell 25: 713–724.
Feng S, Yang Y, Mei Y, Ma L, Zhu DE, Hoti N, Castanares M, Wu M. 2007. Cleavage of RIP3
inactivates its caspase-independent apoptosis pathway by removal of kinase domain. Cell Signal
19: 2056–2067.
Feoktistova M, Geserick P, Kellert B, Dimitrova DP, Langlais C, Hupe M, Cain K, MacFarlane M,
Hacker G, Leverkus M. 2011a. cIAPs block Ripoptosome formation, a RIP1/caspase-8 containing
intracellular cell death complex differentially regulated by cFLIP isoforms. Mol Cell 43: 449–463.
Feoktistova M, Geserick P, Kellert B, Dimitrova DP, Langlais C, Hupe M, Cain K, MacFarlane M,
Häcker G, Leverkus M. 2011b. cIAPs block Ripoptosome formation, a RIP1/caspase-8 containing
intracellular cell death complex differentially regulated by cFLIP isoforms. Mol Cell 43: 449–463.
Fernandez Y, Verhaegen M, Miller TP, Rush JL, Steiner P, Opipari AW Jr, Lowe SW, Soengas MS.
2005. Differential regulation of noxa in normal melanocytes and melanoma cells by proteasome
inhibition: Therapeutic implications. Cancer Res 65: 6294–6304.
Fimia GM, Corazzari M, Antonioli M, Piacentini M. 2012. Ambra1 at the crossroad between
autophagy and cell death. Oncogene 32: 3311–3318.
Fink SL, Cookson BT. 2006. Caspase-1-dependent pore formation during pyroptosis leads to osmotic
lysis of infected host macrophages. Cell Microbiol 8: 1812–1825.
Fox CJ, Hammerman PS, Cinalli RM, Master SR, Chodosh LA, Thompson CB. 2003. The
serine/threonine kinase Pim-2 is a transcriptionally regulated apoptotic inhibitor. Genes Dev 17:
1841–1854.
Franchi L, Eigenbrod T, Munoz-Planillo R, Nunez G. 2009. The inflammasome: A caspase-1-
activation platform that regulates immune responses and disease pathogenesis. Nat Immunol 10:
241–247.
Franchi L, Munoz-Planillo R, Nunez G. 2012. Sensing and reacting to microbes through the
inflammasomes. Nat Immunol 13: 325–332.
Galluzzi L, Maiuri MC, Vitale I, Zischka H, Castedo M, Zitvogel L, Kroemer G. 2007. Cell death
modalities: Classification and pathophysiological implications. Cell Death Differ 14: 1237–1243.
Galluzzi L, Vitale I, Abrams JM, Alnemri ES, Baehrecke EH, Blagosklonny MV, Dawson TM,
Dawson VL, El-Deiry WS, Fulda S, et al. 2012. Molecular definitions of cell death subroutines:
Recommendations of the Nomenclature Committee on Cell Death 2012. Cell Death Differ 19: 107–
120.
Ganley IG, Lam du H, Wang J, Ding X, Chen S, Jiang X. 2009. ULK1. ATG13.FIP200 complex
mediates mTOR signaling and is essential for autophagy. J Biol Chem 284: 12297–12305.
Geserick P, Hupe M, Moulin M, Wong WW, Feoktistova M, Kellert B, Gollnick H, Silke J, Leverkus
M. 2009. Cellular IAPs inhibit a cryptic CD95-induced cell death by limiting RIP1 kinase
recruitment. J Cell Biol 187: 1037–1054.
Gilley J, Coffer PJ, Ham J. 2003. FOXO transcription factors directly activate bim gene expression and
promote apoptosis in sympathetic neurons. J Cell Biol 162: 613–622.
Goldstein JC, Waterhouse NJ, Juin P, Evan GI, Green DR. 2000. The coordinate release of cytochrome
c during apoptosis is rapid, complete and kinetically invariant. Nat Cell Biol 2: 156–162.
Goyal L, McCall K, Agapite J, Hartwieg E, Steller H. 2000. Induction of apoptosis by Drosophila
reaper, hid and grim through inhibition of IAP function. EMBO J 19: 589–597.
Green DR. 2011a. Immunology: A heavyweight knocked out. Nature 479: 48–50.
Green DR. 2011b. Means to an end. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY.
Guo K, Searfoss G, Krolikowski D, Pagnoni M, Franks C, Clark K, Yu KT, Jaye M, Ivashchenko Y.
2001. Hypoxia induces the expression of the pro-apoptotic gene BNIP3. Cell Death Differ 8: 367–
376.
Halestrap AP, Pasdois P. 2009. The role of the mitochondrial permeability transition pore in heart
disease. Biochim Biophys Acta 1787: 1402–1415.
Hanada T, Noda NN, Satomi Y, Ichimura Y, Fujioka Y, Takao T, Inagaki F, Ohsumi Y. 2007. The
Atg12-Atg5 conjugate has a novel E3-like activity for protein lipidation in autophagy. J Biol Chem
282: 37298–37302.
Harada H, Becknell B, Wilm M, Mann M, Huang LJ, Taylor SS, Scott JD, Korsmeyer SJ. 1999.
Phosphorylation and inactivation of BAD by mitochondria-anchored protein kinase A. Mol Cell 3:
413–422.
Harada H, Andersen JS, Mann M, Terada N, Korsmeyer SJ. 2001. p70S6 kinase signals cell survival as
well as growth, inactivating the proapoptotic molecule BAD. Proc Natl Acad Sci 98: 9666–9670.
* Hardie DG. 2012. Organismal carbohydrate and lipid homeostasis. Cold Spring Harb Perspect Biol
4: a006031.
He C, Levine B. 2010. The Beclin 1 interactome. Curr Opin Cell Biol 22: 140–149.
He S, Wang L, Miao L, Wang T, Du F, Zhao L, Wang X. 2009. Receptor interacting protein kinase-3
determines cellular necrotic response to TNF-α. Cell 137: 1100–1111.
He C, Bassik MC, Moresi V, Sun K, Wei Y, Zou Z, An Z, Loh J, Fisher J, Sun Q, et al. 2012. Exercise-
induced BCL2-regulated autophagy is required for muscle glucose homeostasis. Nature 481: 511–
515.
Heibein JA, Goping IS, Barry M, Pinkoski MJ, Shore GC, Green DR, Bleackley RC. 2000. Granzyme
B-mediated cytochrome c release is regulated by the Bcl-2 family members bid and Bax. J Exp Med
192: 1391–1402.
Hershko T, Ginsberg D. 2004. Up-regulation of Bcl-2 homology 3 BH3-only proteins by E2F1
mediates apoptosis. J Biol Chem 279: 8627–8634.
Higuchi M, Tomioka M, Takano J, Shirotani K, Iwata N, Masumoto H, Maki M, Itohara S, Saido TC.
2005. Distinct mechanistic roles of calpain and caspase activation in neurodegeneration as revealed
in mice overexpressing their specific inhibitors. Biol Chem 280: 15229–15237.
Hitomi J, Christofferson DE, Ng A, Yao J, Degterev A, Xavier RJ, Yuan J. 2008. Identification of a
molecular signaling network that regulates a cellular necrotic cell death pathway. Cell 135: 1311–
1323.
Hofmann K, Bucher P, Tschopp J. 1997. The CARD domain: A new apoptotic signalling motif. Trends
Biochem Sci 22: 155–156.
Holler N, Zaru R, Micheau O, Thome M, Attinger A, Valitutti S, Bodmer JL, Schneider P, Seed B,
Tschopp J. 2000. Fas triggers an alternative, caspase-8-independent cell death pathway using the
kinase RIP as effector molecule. Nat Immunol 1: 489–495.
Hornung V, Ablasser A, Charrel-Dennis M, Bauernfeind F, Horvath G, Caffrey DR, Latz E, Fitzgerald
KA. 2009. AIM2 recognizes cytosolic dsDNA and forms a caspase-1-activating inflammasome
with ASC. Nature 458: 514–518.
Hosokawa N, Hara T, Kaizuka T, Kishi C, Takamura A, Miura Y, Iemura S, Natsume T, Takehana K,
Yamada N, et al. 2009. Nutrient-dependent mTORC1 association with the ULK1-Atg13-FIP200
complex required for autophagy. Mol Biol Cell 20: 1981–1991.
Hubner A, Barrett T, Flavell RA, Davis RJ. 2008. Multisite phosphorylation regulates Bim stability and
apoptotic activity. Mol Cell 30: 415–425.
Imtiyaz HZ, Rosenberg S, Zhang Y, Rahman ZSM, Hou Y-J, Manser T, Zhang J. 2006. The Fas-
associated death domain protein (FADD) is required in apoptosis and TLR-induced proliferative
responses in B cells. J Immunol 176: 6852–6861.
Itakura E, Kishi-Itakura C, Mizushima N. 2012. The hairpin-type tail-anchored SNARE syntaxin 17
targets to autophagosomes for fusion with endosomes/lysosomes. Cell 151: 1256–1269.
Janssens S, Tinel A, Lippens S, Tschopp J. 2005. PIDD mediates NF-κB activation in response to DNA
damage. Cell 123: 1079–1092.
Jeffers JR, Parganas E, Lee Y, Yang C, Wang J, Brennan J, MacLean KH, Han J, Chittenden T, Ihle
JN, et al. 2003. Puma is an essential mediator of p53-dependent and -independent apoptotic
pathways. Cancer Cell 4: 321–328.
Jost PJ, Grabow S, Gray D, McKenzie MD, Nachbur U, Huang DC, Bouillet P, Thomas HE, Borner C,
Silke J, et al. 2009. XIAP discriminates between type I and type II FAS-induced apoptosis. Nature
460: 1035–1039.
Jung CH, Jun CB, Ro SH, Kim YM, Otto NM, Cao J, Kundu M, Kim DH. 2009. ULK-Atg13-FIP200
complexes mediate mTOR signaling to the autophagy machinery. Mol Biol Cell 20: 1992–2003.
Kanuka H, Sawamoto K, Inohara N, Matsuno K, Okano H, Miura M. 1999. Control of the cell death
pathway by Dapaf-1, a Drosophila Apaf-1/CED-4-related caspase activator. Mol Cell 4: 757–769.
Kayagaki N, Warming S, Lamkanfi M, Vande Walle L, Louie S, Dong J, Newton K, Qu Y, Liu J,
Heldens S, et al. 2011. Non-canonical inflammasome activation targets caspase 11. Nature 479:
117–121.
Kerr JF, Wyllie AH, Currie AR. 1972. Apoptosis: A basic biological phenomenon with wide-ranging
implications in tissue kinetics. Br J Cancer 26: 239–257.
Kerr JB, Hutt KJ, Michalak EM, Cook M, Vandenberg CJ, Liew SH, Bouillet P, Mills A, Scott CL,
Findlay JK, et al. 2012. DNA damage-induced primordial follicle oocyte apoptosis and loss of
fertility require TAp63-mediated induction of Puma and Noxa. Mol Cell 48: 343–352.
Kersse K, Verspurten J, Vanden Berghe T, Vandenabeele P. 2011. The death-fold superfamily of
homotypic interaction motifs. Trends Biochem Sci 36: 541–552.
Kim J-Y, Ahn H-J, Ryu J-H, Suk K, Park J-H. 2004. BH3-only protein Noxa is a mediator of hypoxic
cell death induced by hypoxia-inducible factor 1α. J Exp Med 199: 113–124.
Kim HE, Du F, Fang M, Wang X. 2005. Formation of apoptosome is initiated by cytochrome c-induced
dATP hydrolysis and subsequent nucleotide exchange on Apaf-1. Proc Natl Acad Sci 102: 17545–
17550.
Kim YS, Morgan MJ, Choksi S, Liu ZG. 2007. TNF-induced activation of the Nox1 NADPH oxidase
and its role in the induction of necrotic cell death. Mol Cell 26: 675–687.
Kim HE, Jiang X, Du F, Wang X. 2008. PHAPI, CAS, and Hsp70 promote apoptosome formation by
preventing Apaf-1 aggregation and enhancing nucleotide exchange on Apaf-1. Mol Cell 30: 239–
247.
Kim J, Kundu M, Viollet B, Guan KL. 2011. AMPK and mTOR regulate autophagy through direct
phosphorylation of Ulk1. Nat Cell Biol 13: 132–141.
Klumpp S, Selke D, Krieglstein J. 2003. Protein phosphatase type 2C dephosphorylates BAD.
Neurochem Int 42: 555–560.
Konishi Y, Lehtinen M, Donovan N, Bonni A. 2002. Cdc2 phosphorylation of BAD links the cell cycle
to the cell death machinery. Mol Cell 9: 1005–1016.
Korsmeyer SJ, Wei MC, Saito M, Weiler S, Oh KJ, Schlesinger PH. 2000. Proapoptotic cascade
activates BID, which oligomerizes BAK or BAX into pores that result in the release of cytochrome
c. Cell Death Differ 7: 1166–1173.
Kothakota S, Azuma T, Reinhard C, Klippel A, Tang J, Chu K, McGarry TJ, Kirschner MW, Koths K,
Kwiatkowski DJ, et al. 1997. Caspase-3–generated fragment of gelsolin: Effector of morphological
change in apoptosis. Science 278: 294–298.
Krajcovic M, Johnson NB, Sun Q, Normand G, Hoover N, Yao E, Richardson AL, King RW, Cibas
ES, Schnitt SJ, et al. 2011. A non-genetic route to aneuploidy in human cancers. Nat Cell Biol 13:
324–330.
Kroemer G, Levine B. 2008. Autophagic cell death: The story of a misnomer. Nat Rev Mol Cell Biol 9:
1004–1010.
Kroemer G, Marino G, Levine B. 2010. Autophagy and the integrated stress response. Mol Cell 40:
280–293.
Krumschnabel G, Manzl C, Villunger A. 2009. Caspase 2: Killer, savior and safeguard—emerging
versatile roles for an ill-defined caspase. Oncogene 28: 3093–3096.
Kuwana T, Mackey MR, Perkins G, Ellisman MH, Latterich M, Schneiter R, Green DR, Newmeyer
DD. 2002. Bid, Bax, and lipids cooperate to form supramolecular openings in the outer
mitochondrial membrane. Cell 111: 331–342.
Kuwana T, Bouchier-Hayes L, Chipuk JE, Bonzon C, Sullivan BA, Green DR, Newmeyer DD. 2005.
BH3 domains of BH3-only proteins differentially regulate Bax-mediated mitochondrial membrane
permeabilization both directly and indirectly. Mol Cell 17: 525–535.
Lamkanfi M, Moreira LO, Makena P, Spierings DC, Boyd K, Murray PJ, Green DR, Kanneganti TD.
2009. Caspase-7 deficiency protects from endotoxin-induced lymphocyte apoptosis and improves
survival. Blood 113: 2742–2745.
* Laplante M, Sabatini DM. 2012. mTOR signaling. Cold Spring Harb Perspect Biol 4: a011593.
Lartigue L, Kushnareva Y, Seong Y, Lin H, Faustin B, Newmeyer DD. 2009. Caspase-independent
mitochondrial cell death results from loss of respiration, not cytotoxic protein release. Mol Biol Cell
20: 4871–4884.
Lee EF, Clarke OB, Evangelista M, Feng Z, Speed TP, Tchoubrieva EB, Strasser A, Kalinna BH,
Colman PM, Fairlie WD. 2011. Discovery and molecular characterization of a Bcl-2-regulated cell
death pathway in schistosomes. Proc Natl Acad Sci 108: 6999–7003.
Lei K, Davis RJ. 2003. JNK phosphorylation of Bim-related members of the Bcl2 family induces Bax-
dependent apoptosis. Proc Natl Acad Sci 100: 2432–2437.
Letai A, Bassik MC, Walensky LD, Sorcinelli MD, Weiler S, Korsmeyer SJ. 2002. Distinct BH3
domains either sensitize or activate mitochondrial apoptosis, serving as prototype cancer
therapeutics. Cancer Cell 2: 183–192.
Levine B, Kroemer G. 2008. Autophagy in the pathogenesis of disease. Cell 132: 27–42.
Ley R, Balmanno K, Hadfield K, Weston C, Cook SJ. 2003. Activation of the ERK1/2 signaling
pathway promotes phosphorylation and proteasome-dependent degradation of the BH3-only
protein, Bim. J Biol Chem 278: 18811–18816.
Li H, Zhu H, Xu CJ, Yuan J. 1998. Cleavage of BID by caspase 8 mediates the mitochondrial damage
in the Fas pathway of apoptosis. Cell 94: 491–501.
Li YM, Wen Y, Zhou BP, Kuo H-P, Ding Q, Hung M-C. 2003. Enhancement of Bik antitumor effect
by Bik mutants. Cancer Res 63: 7630–7633.
* Lim K-H, Staudt LM. 2013. Toll-like receptor signaling. Cold Spring Harb Perspect Biol 5: a011247.
Lin Y, Devin A, Rodriguez Y, Liu ZG. 1999. Cleavage of the death domain kinase RIP by caspase 8
prompts TNF-induced apoptosis. Genes Dev 13: 2514–2526.
Lin Y, Ma W, Benchimol S. 2000. Pidd, a new death-domain-containing protein, is induced by p53 and
promotes apoptosis. Nat Genet 26: 122–127.
Lisi S, Mazzon I, White K. 2000. Diverse domains of THREAD/DIAP1 are required to inhibit
apoptosis induced by REAPER and HID in Drosophila. Genetics 154: 669–678.
Liu L, Feng D, Chen G, Chen M, Zheng Q, Song P, Ma Q, Zhu C, Wang R, Qi W, et al. 2012.
Mitochondrial outer-membrane protein FUNDC1 mediates hypoxia-induced mitophagy in
mammalian cells. Nat Cell Biol 14: 177–185.
Llambi F, Moldoveanu T, Tait SWG, Bouchier-Hayes L, Temirov J, McCormick LL, Dillon CP, Green
DR. 2011. A unified model of mammalian BCL-2 protein family interactions at the mitochondria.
Mol Cell 44: 517–531.
Lu JV, Weist BM, van Raam BJ, Marro BS, Nguyen LV, Srinivas P, Bell BD, Luhrs KA, Lane TE,
Salvesen GS, et al. 2011. Complementary roles of Fas-associated death domain (FADD) and
receptor interacting protein kinase-3 (RIPK3) in T-cell homeostasis and antiviral immunity. Proc
Natl Acad Sci 108: 15312–15317.
Luo X, Budihardjo I, Zou H, Slaughter C, Wang X. 1998. Bid, a Bcl2 interacting protein, mediates
cytochrome c release from mitochondria in response to activation of cell surface death receptors.
Cell 94: 481–490.
Ma C, Ying C, Yuan Z, Song B, Li D, Liu Y, Lai B, Li W, Chen R, Ching Y-P, et al. 2007. dp5/HRK is
a c-Jun target gene and required for apoptosis induced by potassium deprivation in cerebellar
granule neurons. J Biol Chem 282: 30901–30909.
Maiuri MC, Criollo A, Tasdemir E, Vicencio JM, Tajeddine N, Hickman JA, Geneste O, Kroemer G.
2007a. BH3-only proteins and BH3 mimetics induce autophagy by competitively disrupting the
interaction between Beclin 1 and Bcl-2/Bcl-XL. Autophagy 3: 374–376.
Maiuri MC, Le Toumelin G, Criollo A, Rain JC, Gautier F, Juin P, Tasdemir E, Pierron G, Troulinaki
K, Tavernarakis N, et al. 2007b. Functional and physical interaction between Bcl-XL and a BH3-
like domain in Beclin-1. EMBO J 26: 2527–2539.
Malladi S, Challa-Malladi M, Fearnhead HO, Bratton SB. 2009. The Apaf-1•procaspase-9 apoptosome
complex functions as a proteolytic-based molecular timer. EMBO J 28: 1916–1925.
Mandic A, Viktorsson K, Strandberg L, Heiden T, Hansson J, Linder S, Shoshan MC. 2002. Calpain-
mediated Bid cleavage and calpain-independent Bak modulation: Two separate pathways in
cisplatin-induced apoptosis. Mol Cell Biol 22: 3003–3013.
Manzl C, Krumschnabel G, Bock F, Sohm B, Labi V, Baumgartner F, Logette E, Tschopp J, Villunger
A. 2009. Caspase-2 activation in the absence of PIDDosome formation. J Cell Biol 185: 291–303.
Manzl C, Peintner L, Krumschnabel G, Bock F, Labi V, Drach M, Newbold A, Johnstone R, Villunger
A. 2012. PIDDosome-independent tumor suppression by Caspase 2. Cell Death Differ 19: 1722–
1732.
Martinez J, Almendinger J, Oberst A, Ness R, Dillon CP, Fitzgerald P, Hengartner MO, Green DR.
2011. Microtubule-associated protein 1 light chain 3α (LC3)-associated phagocytosis is required for
the efficient clearance of dead cells. Proc Natl Acad Sci 108: 17396–17401.
Martinon F, Burns K, Tschopp J. 2002. The inflammasome: A molecular platform triggering activation
of inflammatory caspases and processing of proIL-β. Mol Cell 10: 417–426.
Matsui H, Asou H, Inaba T. 2007. Cytokines direct the regulation of Bim mRNA stability by heat-
shock cognate protein 70. Mol Cell 25: 99–112.
Matsumura H, Shimizu Y, Ohsawa Y, Kawahara A, Uchiyama Y, Nagata S. 2000. Necrotic death
pathway in Fas receptor signaling. J Cell Biol 151: 1247–1256.
Maurer U, Charvet C, Wagman AS, Dejardin E, Green DR. 2006. Glycogen synthase kinase-3
regulates mitochondrial outer membrane permeabilization and apoptosis by destabilization of MCL-
1. Mol Cell 21: 749–760.
McStay GP, Salvesen GS, Green DR. 2008. Overlapping cleavage motif selectivity of caspases:
Implications for analysis of apoptotic pathways. Cell Death Differ 15: 322–331.
Mei Y, Yong J, Liu H, Shi Y, Meinkoth J, Dreyfuss G, Yang X. 2010. tRNA binds to cytochrome c and
inhibits caspase activation. Mol Cell 37: 668–678.
Micheau O, Lens S, Gaide O, Alevizopoulos K, Tschopp J. 2001. NF-κB signals induce the expression
of c-FLIP. Mol Cell Biol 21: 5299–5305.
Micheau O, Thome M, Schneider P, Holler N, Tschopp J, Nicholson DW, Briand C, Grutter MG. 2002.
The long form of FLIP is an activator of caspase 8 at the Fas death-inducing signaling complex. J
Biol Chem 277: 45162–45171.
Moulin M, Anderton H, Voss AK, Thomas T, Wong WW, Bankovacki A, Feltham R, Chau D, Cook
WD, Silke J, et al. 2012. IAPs limit activation of RIP kinases by TNF receptor 1 during
development. EMBO J 31: 1679–1691.
Munoz-Pinedo C, Guio-Carrion A, Goldstein JC, Fitzgerald P, Newmeyer DD, Green DR. 2006.
Different mitochondrial intermembrane space proteins are released during apoptosis in a manner
that is coordinately initiated but can vary in duration. Proc Natl Acad Sci 103: 11573–11578.
Naik E, Michalak EM, Villunger A, Adams JM, Strasser A. 2007. Ultraviolet radiation triggers
apoptosis of fibroblasts and skin keratinocytes mainly via the BH3-only protein Noxa. J Cell Biol
176: 415–424.
Nakagawa T, Shimizu S, Watanabe T, Yamaguchi O, Otsu K, Yamagata H, Inohara H, Kubo T,
Tsujimoto Y. 2005. Cyclophilin D-dependent mitochondrial permeability transition regulates some
necrotic but not apoptotic cell death. Nature 434: 652–658.
Nakano K, Vousden KH. 2001. PUMA, a novel proapoptotic gene, is induced by p53. Mol Cell 7: 683–
694.
Narendra DP, Youle RJ. 2011. Targeting mitochondrial dysfunction: Role for PINK1 and Parkin in
mitochondrial quality control. Antioxid Redox Signal 14: 1929–1938.
Newmeyer DD, Bossy-Wetzel E, Kluck RM, Wolf BB, Beere HM, Green DR. 2000. Bcl-xL does not
inhibit the function of Apaf-1. Cell Death Differ 7: 402–407.
* Newton K, Dixit VM. 2012. Signaling in innate immunity and inflammation. Cold Spring Harb
Perspect Biol 4: a006049.
Nikiforov MA, Riblett M, Tang W-H, Gratchouck V, Zhuang D, Fernandez Y, Verhaegen M,
Varambally S, Chinnaiyan AM, Jakubowiak AJ, et al. 2007. Tumor cell-selective regulation of
NOXA by c-MYC in response to proteasome inhibition. Proc Natl Acad Sci 104: 19488–19493.
Nikrad M, Johnson T, Puthalalath H, Coultas L, Adams J, Kraft AS. 2005. The proteasome inhibitor
bortezomib sensitizes cells to killing by death receptor ligand TRAIL via BH3–only proteins Bik
and Bim. Mol Cancer Ther 4: 443–449.
Nutt LK, Margolis SS, Jensen M, Herman CE, Dunphy WG, Rathmell JC, Kornbluth S. 2005.
Metabolic regulation of oocyte cell death through the CaMKII-mediated phosphorylation of caspase
2. Cell 123: 89–103.
Nutt LK, Buchakjian MR, Gan E, Darbandi R, Yoon SY, Wu JQ, Miyamoto YJ, Gibbons JA, Andersen
JL, Freel CD, et al. 2009. Metabolic control of oocyte apoptosis mediated by 14-3-3ζ-regulated
dephosphorylation of caspase 2. Dev Cell 16: 856–866.
Oberst A, Pop C, Tremblay AG, Blais V, Denault J-B, Salvesen GS, Green DR. 2010. Inducible
dimerization and inducible cleavage reveal a requirement for both processes in caspase-8 activation.
J Biol Chem 285: 16632–16642.
Oberst A, Dillon CP, Weinlich R, McCormick LL, Fitzgerald P, Pop C, Hakem R, Salvesen GS, Green
DR. 2011. Catalytic activity of the caspase-8-FLIPL complex inhibits RIPK3-dependent necrosis.
Nature 471: 363–367.
Oberstein A, Jeffrey PD, Shi Y. 2007. Crystal structure of the Bcl-XL-Beclin 1 peptide complex:
Beclin 1 is a novel BH3-only protein. J Biol Chem 282: 13123–13132.
Oda E, Ohki R, Murasawa H, Nemoto J, Shibue T, Yamashita T, Tokino T, Taniguchi T, Tanaka N.
2000. Noxa, a BH3-only member of the Bcl-2 family and candidate mediator of p53-induced
apoptosis. Science 288: 1053–1058.
O’Donnell MA, Perez-Jimenez E, Oberst A, Ng A, Massoumi R, Xavier R, Green DR, Ting AT. 2011.
Caspase 8 inhibits programmed necrosis by processing CYLD. Nat Cell Biol 13: 1437–1442.
Overholtzer M, Mailleux AA, Mouneimne G, Normand G, Schnitt SJ, King RW, Cibas ES, Brugge JS.
2007. A nonapoptotic cell death process, entosis, that occurs by cell-in-cell invasion. Cell 131:
966–979.
Pattingre S, Tassa A, Qu X, Garuti R, Liang XH, Mizushima N, Packer M, Schneider MD, Levine B.
2005. Bcl-2 antiapoptotic proteins inhibit Beclin 1-dependent autophagy. Cell 122: 927–939.
Ploner C, Rainer J, Niederegger H, Eduardoff M, Villunger A, Geley S, Kofler R. 2008. The BCL2
rheostat in glucocorticoid-induced apoptosis of acute lymphoblastic leukemia. Leukemia 22: 370–
377.
Pop C, Fitzgerald P, Green DR, Salvesen GS. 2007. Role of proteolysis in caspase-8 activation and
stabilization. Biochemistry 46: 4398–4407.
Pop C, Oberst A, Drag M, Van Raam BJ, Riedl SJ, Green DR, Salvesen GS. 2011. FLIPL induces
caspase-8 activity in the absence of interdomain caspase-8 cleavage and alters substrate specificity.
Biochem J 433: 447–457.
Putcha GV, Le S, Frank S, Besirli CG, Clark K, Chu B, Alix S, Youle RJ, LaMarche A, Maroney AC,
et al. 2003. JNK-mediated BIM phosphorylation potentiates BAX-dependent apoptosis. Neuron 38:
899–914.
Puthalakath H, O’Reilly LA, Gunn P, Lee L, Kelly PN, Huntington ND, Hughes PD, Michalak EM,
McKimm-Breschkin J, Motoyama N, et al. 2007. ER stress triggers apoptosis by activating BH3-
only protein Bim. Cell 129: 1337–1349.
Pyati UJ, Gjini E, Carbonneau S, Lee JS, Guo F, Jette CA, Kelsell DP, Look AT. 2011. p63 mediates
an apoptotic response to pharmacological and disease-related ER stress in the developing epidermis.
Dev Cell 21: 492–505.
Qin JZ, Ziffra J, Stennett L, Bodner B, Bonish BK, Chaturvedi V, Bennett F, Pollock PM, Trent JM,
Hendrix MJ, et al. 2005. Proteasome inhibitors trigger NOXA-mediated apoptosis in melanoma and
myeloma cells. Cancer Res 65: 6282–6293.
Ramjaun AR, Tomlinson S, Eddaoudi A, Downward J. 2007. Upregulation of two BH3–only proteins,
Bmf and Bim, during TGF β-induced apoptosis. Oncogene 26: 970–981.
Real PJ, Sanz C, Gutierrez O, Pipaon C, Zubiaga AM, Fernandez-Luna JL. 2006. Transcriptional
activation of the proapoptotic bik gene by E2F proteins in cancer cells. FEBS Lett 580: 5905–5909.
Reiners JJ, Caruso JA, Mathieu P, Chelladurai B, Yin X-M, Kessel D. 2002. Release of cytochrome c
and activation of procaspase 9 following lysosomal photodamage involves Bid cleavage. Cell
Death Differ 9: 934–944.
* Rhind N, Russel P. 2012. Signaling pathways that regulate cell division. Cold Spring Harb Perspect
Biol 4: a005942.
Ribe EM, Jean YY, Goldstein RL, Manzl C, Stefanis L, Villunger A, Troy CM. 2012. Neuronal
caspase-2 activity and function requires RAIDD, but not PIDD. Biochem J 444: 591–599.
Ricchelli F, Sileikyte J, Bernardi P. 2011. Shedding light on the mitochondrial permeability transition.
Biochim Biophys Acta 1807: 482–490.
Ricci JE, Munoz-Pinedo C, Fitzgerald P, Bailly-Maitre B, Perkins GA, Yadava N, Scheffler IE,
Ellisman MH, Green DR. 2004. Disruption of mitochondrial function during apoptosis is mediated
by caspase cleavage of the p75 subunit of complex I of the electron transport chain. Cell 117: 773–
786.
Rodriguez J, Lazebnik Y. 1999. Caspase 9 and APAF-1 form an active holoenzyme. Genes Dev 13:
3179–3184.
Rodriguez A, Oliver H, Zou H, Chen P, Wang X, Abrams JM. 1999. Dark is a Drosophila homologue
of Apaf-1/CED-4 and functions in an evolutionarily conserved death pathway. Nat Cell Biol 1:
272–279.
Rodriguez A, Chen P, Oliver H, Abrams JM. 2002. Unrestrained caspase-dependent cell death caused
by loss of Diap1 function requires the Drosophila Apaf-1 homolog, Dark. EMBO J 21: 2189–2197.
Rudel T, Bokoch GM. 1997. Membrane and morphological changes in apoptotic cells regulated by
caspase-mediated activation of PAK2. Science 276: 1571–1574.
Sakahira H, Enari M, Nagata S. 1998. Cleavage of CAD inhibitor in CAD activation and DNA
degradation during apoptosis. Nature 391: 96–99.
Salvesen GS, Riedl SJ. 2008. Caspase mechanisms. Adv Exp Med Biol 615: 13–23.
Sandoval H, Thiagarajan P, Dasgupta SK, Schumacher A, Prchal JT, Chen M, Wang J. 2008. Essential
role for Nix in autophagic maturation of erythroid cells. Nature 454: 232–235.
Sanjuan MA, Dillon CP, Tait SWG, Moshiach S, Dorsey F, Connell S, Komatsu M, Tanaka K,
Cleveland JL, Withoff S, et al. 2007. Toll-like receptor signalling in macrophages links the
autophagy pathway to phagocytosis. Nature 450: 1253–1257.
Sanz C, Mellstrom B, Link WA, Naranjo JR, Fernandez-Luna JL. 2001. Interleukin 3-dependent
activation of DREAM is involved in transcriptional silencing of the apoptotic Hrk gene in
hematopoietic progenitor cells. EMBO J 20: 2286–2292.
Schinzel AC, Takeuchi O, Huang Z, Fisher JK, Zhou Z, Rubens J, Hetz C, Danial NN, Moskowitz MA,
Korsmeyer SJ. 2005. Cyclophilin D is a component of mitochondrial permeability transition and
mediates neuronal cell death after focal cerebral ischemia. Proc Natl Acad Sci 102: 12005–12010.
Schulze-Osthoff K, Krammer PH, Droge W. 1994. Divergent signalling via APO-1/Fas and the TNF
receptor, two homologous molecules involved in physiological cell death. EMBO J 13: 4587–4596.
Schürmann A, Mooney AF, Sanders LC, Sells MA, Wang HG, Reed JC, Bokoch GM. 2000. p21-
Activated kinase 1 phosphorylates the death agonist bad and protects cells from apoptosis. Mol Cell
Biol 20: 453–461.
Schweers RL, Zhang J, Randall MS, Loyd MR, Li W, Dorsey FC, Kundu M, Opferman JT, Cleveland
JL, Miller JL, et al. 2007. NIX is required for programmed mitochondrial clearance during
reticulocyte maturation. Proc Natl Acad Sci 104: 19500–19505.
Schwickart M, Huang X, Lill JR, Liu J, Ferrando R, French DM, Maecker H, O’Rourke K, Bazan F,
Eastham-Anderson J, et al. 2010. Deubiquitinase USP9X stabilizes MCL1 and promotes tumour
cell survival. Nature 463: 103–107.
Sebbagh M, Renvoizé C, Hamelin J, Riché N, Bertoglio J, Bréard J. 2001. Caspase-3-mediated
cleavage of ROCK I induces MLC phosphorylation and apoptotic membrane blebbing. Nature Cell
Biol 3: 346–352.
Seshagiri S, Miller LK. 1997. Caenorhabditis elegans CED-4 stimulates CED-3 processing and CED-
3-induced apoptosis. Curr Biol 7: 455–460.
* Sever R, Brugge JS. 2014. Signal transduction in cancer. Cold Spring Harb Perspect Med doi:
10.1101/cshperspect.a006098
Shaid S, Brandts CH, Serve H, Dikic I. 2013. Ubiquitination and selective autophagy. Cell Death Differ
20: 21–30.
Shaw J, Zhang T, Rzeszutek M, Yurkova N, Baetz D, Davie JR, Kirshenbaum LA. 2006.
Transcriptional silencing of the death gene BNIP3 by cooperative action of NF-κB and histone
deacetylase 1 in ventricular myocytes. Circ Res 99: 1347–1354.
Shaw J, Yurkova N, Zhang T, Gang H, Aguilar F, Weidman D, Scramstad C, Weisman H,
Kirshenbaum LA. 2008. Antagonism of E2F-1 regulated Bnip3 transcription by NF-κB is essential
for basal cell survival. Proc Natl Acad Sci 105: 20734–20739.
Shen S, Kepp O, Kroemer G. 2012. The end of autophagic cell death? Autophagy 8: 1–3.
Slee EA, Harte MT, Kluck RM, Wolf BB, Casiano CA, Newmeyer DD, Wang HG, Reed JC,
Nicholson DW, Alnemri ES, et al. 1999. Ordering the cytochrome-c-initiated caspase cascade:
Hierarchical activation of caspases-2, -3, -6, -7, -8, and -10 in a caspase-9-dependent manner. J Cell
Biol 144: 281–292.
Sowter HM, Ratcliffe PJ, Watson P, Greenberg AH, Harris AL. 2001. HIF-1-dependent regulation of
hypoxic induction of the cell death factors BNIP3 and NIX in human tumors. Cancer Res 61:
6669–6673.
Spector MS, Desnoyers S, Hoeppner DJ, Hengartner MO. 1997. Interaction between the C. elegans
cell-death regulators CED-9 and CED-4. Nature 385: 653–656.
Spender LC, O'Brien DI, Simpson D, Dutt D, Gregory CD, Allday MJ, Clark LJ, Inman GJ. 2009.
TGF-β induces apoptosis in human B cells by transcriptional regulation of BIK and Bcl-xL. Cell
Death Differ 16: 593–602.
Stennicke HR, Deveraux QL, Humke EW, Reed JC, Dixit VM, Salvesen GS. 1999. Caspase 9 can be
activated without proteolytic processing. J Biol Chem 274: 8359–8362.
Stoka V, Turk B, Schendel SL, Kim TH, Cirman T, Snipas SJ, Ellerby LM, Bredesen D, Freeze H,
Abrahamson M, et al. 2001. Lysosomal protease pathways to apoptosis. Cleavage of bid, not
procaspases, is the most likely route. J Biol Chem 276: 3149–3157.
Sun L, Wang H, Wang Z, He S, Chen S, Liao D, Wang L, Yan J, Liu W, Lei X, et al. 2012. Mixed
lineage kinase domain-like protein mediates necrosis signaling downstream of RIP3 kinase. Cell
148: 213–227.
Sutton VR, Davis JE, Cancilla M, Johnstone RW, Ruefli AA, Sedelies K, Browne KA, Trapani JA.
2000. Initiation of apoptosis by granzyme B requires direct cleavage of bid, but not direct granzyme
B-mediated caspase activation. J Exp Med 192: 1403–1414.
Szydlowska K, Tymianski M. 2010. Calcium, ischemia and excitotoxicity. Cell Calcium 47: 122–129.
Tait SW, Green DR. 2010. Mitochondria and cell death: Outer membrane permeabilization and beyond.
Nat Rev Mol Cell Biol 11: 621–632.
Tait SWG, Parsons MJ, Llambi F, Bouchier-Hayes L, Connell S, Muñoz-Pinedo C, Green DR. 2010.
Resistance to caspase-independent cell death requires persistence of intact mitochondria. Dev Cell
18: 802–813.
Tanaka Y, Guhde G, Suter A, Eskelinen EL, Hartmann D, Lullmann-Rauch R, Janssen PM, Blanz J,
von Figura K, Saftig P. 2000. Accumulation of autophagic vacuoles and cardiomyopathy in LAMP-
2-deficient mice. Nature 406: 902–906.
Telliez JB, Bean KM, Lin LL. 2000. LRDD, a novel leucine rich repeat and death domain containing
protein. Biochim Biophys Acta 1478: 280–288.
Tenev T, Bianchi K, Darding M, Broemer M, Langlais C, Wallberg F, Zachariou A, Lopez J,
MacFarlane M, Cain K, et al. 2011. The Ripoptosome, a signaling platform that assembles in
response to genotoxic stress and loss of IAPs. Mol Cell 43: 432–448.
Tinel A, Tschopp J. 2004. The PIDDosome, a protein complex implicated in activation of caspase 2 in
response to genotoxic stress. Science 304: 843–846.
Tinel A, Janssens S, Lippens S, Cuenin S, Logette E, Jaccard B, Quadroni M, Tschopp J. 2007.
Autoproteolysis of PIDD marks the bifurcation between prodeath caspase 2 and prosurvival NF-κB
pathway. EMBO J 26: 197–208.
Troy CM, Stefanis L, Greene LA, Shelanski ML. 1997. Nedd2 is required for apoptosis after trophic
factor withdrawal, but not superoxide dismutase (SOD1) downregulation, in sympathetic neurons
and PC12 cells. J Neurosci 17: 1911–1918.
Troy CM, Rabacchi SA, Friedman WJ, Frappier TF, Brown K, Shelanski ML. 2000. Caspase-2
mediates neuronal cell death induced by β-amyloid. J Neurosci 20: 1386–1392.
Troy CM, Rabacchi SA, Hohl JB, Angelastro JM, Greene LA, Shelanski ML. 2001. Death in the
balance: Alternative participation of the caspase-2 and -9 pathways in neuronal death induced by
nerve growth factor deprivation. J Neurosci 21: 5007–5016.
Upton JP, Austgen K, Nishino M, Coakley KM, Hagen A, Han D, Papa FR, Oakes SA. 2008. Caspase-
2 cleavage of BID is a critical apoptotic signal downstream of endoplasmic reticulum stress. Mol
Cell Biol 28: 3943–3951.
Upton JW, Kaiser WJ, Mocarski ES. 2010. Virus inhibition of RIP3-dependent necrosis. Cell Host
Microbe 7: 302–313.
Upton JW, Kaiser WJ, Mocarski ES. 2012. DAI/ZBP1/DLM-1 complexes with RIP3 to mediate virus-
induced programmed necrosis that is targeted by murine cytomegalovirus vIRA. Cell Host Microbe
11: 290–297.
Urbich C, Knau A, Fichtlscherer S, Walter DH, Brühl T, Potente M, Hofmann WK, de Vos S, Zeiher
AM, Dimmeler S. 2005. FOXO-dependent expression of the proapoptotic protein Bim: Pivotal role
for apoptosis signaling in endothelial progenitor cells. FASEB J 19: 974–976.
Vanden Berghe T, Declercq W, Vandenabeele P. 2007. NADPH oxidases: New players in TNF-
induced necrotic cell death. Mol Cell 26: 769–771.
Vercammen D, Beyaert R, Denecker G, Goossens V, Van Loo G, Declercq W, Grooten J, Fiers W,
Vandenabeele P. 1998. Inhibition of caspases increases the sensitivity of L929 cells to necrosis
mediated by tumor necrosis factor. J Exp Med 187: 1477–1485.
Verma S, Zhao LJ, Chinnadurai G. 2001. Phosphorylation of the pro-apoptotic protein BIK: Mapping
of phosphorylation sites and effect on apoptosis. J Biol Chem 276: 4671–4676.
Villunger A, Michalak EM, Coultas L, Mullauer F, Bock G, Ausserlechner MJ, Adams JM, Strasser A.
2003. p53- and drug-induced apoptotic responses mediated by BH3-only proteins puma and noxa.
Science 302: 1036–1038.
Vosler PS, Brennan CS, Chen J. 2008. Calpain-mediated signaling mechanisms in neuronal injury and
neurodegeneration. Mol Neurobiol 38: 78–100.
Wang Y, Qin ZH. 2010. Molecular and cellular mechanisms of excitotoxic neuronal death. Apoptosis
15: 1382–1402.
Wang HG, Pathan N, Ethell IM, Krajewski S, Yamaguchi Y, Shibasaki F, McKeon F, Bobo T, Franke
TF, Reed JC. 1999. Ca2+-induced apoptosis through calcineurin dephosphorylation of BAD.
Science 284: 339–343.
Wang Y, Guan X, Fok KL, Li S, Zhang X, Miao S, Zong S, Koide SS, Chan HC, Wang L. 2008a. A
novel member of the Rhomboid family, RHBDD1, regulates BIK-mediated apoptosis. Cell Mol Life
Sci 65: 3822–3829.
Wang L, Du F, Wang X. 2008b. TNF-α induces two distinct caspase-8 activation pathways. Cell 133:
693–703.
Wang RC, Wei Y, An Z, Zou Z, Xiao G, Bhagat G, White M, Reichelt J, Levine B. 2012. Akt-
mediated regulation of autophagy and tumorigenesis through Beclin 1 phosphorylation. Science
338: 956–959.
Warr MR, Acoca S, Liu Z, Germain M, Watson M, Blanchette M, Wing SS, Shore GC. 2005. BH3-
ligand regulates access of MCL-1 to its E3 ligase. FEBS Lett 579: 5603–5608.
Wei MC, Lindsten T, Mootha VK, Weiler S, Gross A, Ashiya M, Thompson CB, Korsmeyer SJ. 2000.
tBID, a membrane-targeted death ligand, oligomerizes BAK to release cytochrome c. Genes Dev
14: 2060–2071.
Wei Y, Pattingre S, Sinha S, Bassik M, Levine B. 2008. JNK1-mediated phosphorylation of Bcl-2
regulates starvation-induced autophagy. Mol Cell 30: 678–688.
Wertz IE, Kusam S, Lam C, Okamoto T, Sandoval W, Anderson DJ, Helgason E, Ernst JA, Eby M, Liu
J, et al. 2011. Sensitivity to antitubulin chemotherapeutics is regulated by MCL1 and FBW7.
Nature 471: 110–114.
Wu D, Wallen HD, Inohara N, Nunez G. 1997a. Interaction and regulation of the Caenorhabditis
elegans death protease CED-3 by CED-4 and CED-9. J Biol Chem 272: 21449–21454.
Wu D, Wallen HD, Nunez G. 1997b. Interaction and regulation of subcellular localization of CED-4 by
CED-9. Science 275: 1126–1129.
Xiao C, Srinivasan L, Calado DP, Patterson HC, Zhang B, Wang J, Henderson JM, Kutok JL,
Rajewsky K. 2008. Lymphoproliferative disease and autoimmunity in mice with increased miR-17-
92 expression in lymphocytes. Nat Immunol 9: 405–414.
Yan B, Zemskova M, Holder S, Chin V, Kraft A, Koskinen PJ, Lilly M. 2003. The PIM-2 kinase
phosphorylates BAD on serine 112 and reverses BAD-induced cell death. J Biol Chem 278: 45358–
45367.
Yan N, Gu L, Kokel D, Chai J, Li W, Han A, Chen L, Xue D, Shi Y. 2004. Structural, biochemical, and
functional analyses of CED-9 recognition by the proapoptotic proteins EGL-1 and CED-4. Mol Cell
15: 999–1006.
Yan J, Xiang J, Lin Y, Ma J, Zhang J, Zhang H, Sun J, Danial NN, Liu J, Lin A. 2013. Inactivation of
BAD by IKK inhibits TNFα-induced apoptosis independently of NF-κB activation. Cell 152: 304–
315.
Yang CS, Thomenius MJ, Gan EC, Tang W, Freel CD, Merritt TJ, Nutt LK, Kornbluth S. 2010.
Metabolic regulation of Drosophila apoptosis through inhibitory phosphorylation of Dronc. EMBO
J 29: 3196–3207.
Yoo SJ, Huh JR, Muro I, Yu H, Wang L, Wang SL, Feldman RM, Clem RJ, Muller HA, Hay BA.
2002. Hid, Rpr and Grim negatively regulate DIAP1 levels through distinct mechanisms. Nat Cell
Biol 4: 416–424.
You H, Pellegrini M, Tsuchihara K, Yamamoto K, Hacker G, Erlacher M, Villunger A, Mak TW.
2006. FOXO3a-dependent regulation of Puma in response to cytokine/growth factor withdrawal. J
Exp Med 203: 1657–1663.
Yu J, Zhang L, Hwang PM, Kinzler KW, Vogelstein B. 2001. PUMA induces the rapid apoptosis of
colorectal cancer cells. Mol Cell 7: 673–682.
Yu J, Wang Z, Kinzler KW, Vogelstein B, Zhang L. 2003. PUMA mediates the apoptotic response to
p53 in colorectal cancer cells. Proc Natl Acad Sci 100: 1931–1936.
Yu X, Acehan D, Menetret JF, Booth CR, Ludtke SJ, Riedl SJ, Shi Y, Wang X, Akey CW. 2005. A
structure of the human apoptosome at 12.8 A resolution provides insights into this cell death
platform. Structure 13: 1725–1735.
Yu X, Wang L, Acehan D, Wang X, Akey CW. 2006. Three-dimensional structure of a double
apoptosome formed by the Drosophila Apaf-1 related killer. J Mol Biol 355: 577–589.
Yuan J, Kroemer G. 2010. Alternative cell death mechanisms in development and beyond. Genes Dev
24: 2592–2602.
Yuan J, Shaham S, Ledoux S, Ellis HM, Horvitz HR. 1993. The C. elegans cell death gene ced-3
encodes a protein similar to mammalian interleukin-1 β-converting enzyme. Cell 75: 641–652.
Yuan S, Yu X, Topf M, Ludtke SJ, Wang X, Akey CW. 2010. Structure of an apoptosome-procaspase-
9 CARD complex. Structure 18: 571–583.
Yurkova N, Shaw J, Blackie K, Weidman D, Jayas R, Flynn B, Kirshenbaum LA. 2008. The cell cycle
factor E2F-1 activates Bnip3 and the intrinsic death pathway in ventricular myocytes. Circ Res 102:
472–479.
Zalckvar E, Berissi H, Eisenstein M, Kimchi A. 2009a. Phosphorylation of Beclin 1 by DAP-kinase
promotes autophagy by weakening its interactions with Bcl-2 and Bcl-XL. Autophagy 5: 720–722.
Zalckvar E, Berissi H, Mizrachy L, Idelchuk Y, Koren I, Eisenstein M, Sabanay H, Pinkas-Kramarski
R, Kimchi A. 2009b. DAP-kinase-mediated phosphorylation on the BH3 domain of beclin 1
promotes dissociation of beclin 1 from Bcl-XL and induction of autophagy. EMBO Rep 10: 285–
292.
Zha J, Harada H, Yang E, Jockel J, Korsmeyer SJ. 1996. Serine phosphorylation of death agonist BAD
in response to survival factor results in binding to 14-3-3 not BCL-XL. Cell 87: 619–628.
Zhang DW, Shao J, Lin J, Zhang N, Lu BJ, Lin SC, Dong MQ, Han J. 2009. RIP3, an energy
metabolism regulator that switches TNF-induced cell death from apoptosis to necrosis. Science 325:
332–336.
Zhao J, Jitkaew S, Cai Z, Choksi S, Li Q, Luo J, Liu Z-G. 2012. Mixed lineage kinase domain-like is a
key receptor interacting protein 3 downstream component of TNF-induced necrosis. Proc Natl
Acad Sci 109: 5322–5327.
Zhou L, Song Z, Tittel J, Steller H. 1999. HAC-1, a Drosophila homolog of APAF-1 and CED-4
functions in developmental and radiation-induced apoptosis. Mol Cell 4: 745–755.
Zhu H, Zhang L, Dong F, Guo W, Wu S, Teraishi F, Davis JJ, Chiao PJ, Fang B. 2005. Bik/NBK
accumulation correlates with apoptosis-induction by bortezomib PS-341, Velcade and other
proteasome inhibitors. Oncogene 24: 4993–4999.
Zou H, Henzel WJ, Liu X, Lutschg A, Wang X. 1997. Apaf-1, a human protein homologous to C.
elegans CED-4, participates in cytochrome-c-dependent activation of caspase 3. Cell 90: 405–413.
1Two other processes, microautophagy and chaperone-mediated autophagy, also exist, but are not
further discussed here.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a006080
CHAPTER 20
SUMMARY
Outline
1 Introduction
2 Corruption of MAPK signaling
3 Manipulation of G-protein signaling
4 Highjacking lipid signaling
5 Other actin regulators targeted by pathogens
6 Targeting ubiquitin-mediated signal transduction
7 Conclusion
References
1 INTRODUCTION
Viruses and bacteria both produce proteins that hijack cellular signaling
machinery to ensure their survival despite the constant negative pressure of
their hosts’ innate and adaptive immune systems. Studies of these have not
only revealed methods by which microbial pathogens cause infectious
disease, but also provided critical insights into mechanisms of cellular
regulation, particularly in the case of viral oncoproteins (see Box 1). They
also yielded valuable tools for dissecting mechanisms involved in regulating
signaling.
Bacteria produce different types of virulence factors. Note that these are
not always proteins—they may also be peptides or small molecules—but here
we confine our discussion to proteins. The first type, called a toxin, is
secreted by the pathogen at high concentration and delivered into the host
cytoplasm by a variety of mechanisms, including endocytosis or via protein
pores formed by the bacterial toxin itself (Fig. 1) (Henkel et al. 2010).
Cholera toxin from Vibrio cholerae, for example, is an enzyme that enters the
cell by endocytosis and ADP-ribosylates the αs subunit of the heterotrimeric
Gs protein. This modification prevents αs from hydrolyzing GTP, locking it
into an active state that constitutively stimulates its downstream effector,
adenylyl cyclase (see p. 99 [Sassone-Corsi 2012]). In the intestinal
epithelium, the elevated levels of cyclic AMP (cAMP) generated cause an
efflux of chloride ions and water, resulting in severe diarrhea and
dehydration. Pertussis toxin produced by Bordetella pertussis, which causes
whooping cough, by contrast, ADP ribosylates the αi subunit of the
heterotrimeric Gi protein so that it remains in the GDP-bound state and
cannot be activated by upstream signals. The modified GDP-bound αi is no
longer able to bind to adenylyl cyclase and inhibit production of cAMP.
Pertussis toxin causes major trauma in airways during infection because
signaling pathways are constitutively activated and results in abnormally high
levels of insulin and histamine sensitivity. These toxins have proven very
useful for the elucidation of molecular signaling by G-protein-coupled
receptors (GPCRs) and their downstream signaling partners.
Figure 1. Bacterial secretion systems. Bacteria use several different mechanisms to secrete molecules
into the extracellular space and to translocate molecules into a host cell. Toxins, including peptides and
proteins, are typically secreted through a type I or II secretion system. Effectors are translocated
through a type III or IV secretion system, whereas DNA is only transferred through the type IV
secretion system.
Figure 3. Catalytic triads: same chemistry, different substrates. Yersinia YopJ is proposed to use a
ping-pong mechanism whereby its catalytic cysteine attacks acetyl-CoA to form a covalent acetyl-
enzyme intermediate, followed by a subsequent attack by its second substrate, a hydroxyl on MAPKK,
to transfer the acetyl group to MAPKK. A cysteine protease uses the same mechanism with a peptide to
form a covalent acetyl-enzyme intermediate. The second substrate in this reaction is water and leads to
cleavage of a peptide bond.
During an infection, signals for host survival and apoptosis are induced.
In the presence of YopJ, the default pathway will always be death because
YopJ blocks the NF-κB survival pathway. By inhibiting signaling pathways
that alert the immune system and induce survival signals, YopJ attenuates the
immune response to Yersinia during infection. In contrast, VopA/P, a YopJ
relative from the seafood-borne pathogen Vibrio parahaemolyticus that
causes food poisoning, inhibits MAPK signaling pathways but not the NF-κB
pathway (Fig. 2) (Trosky et al. 2004). Additionally, VopA/P acetylates not
only the activation loop of MAPKKs but also a conserved lysine residue in
the catalytic loop of MAPKKs that is required for coordination of the γ-
phosphate of ATP (Trosky et al. 2007). This inhibits the binding of ATP but
not ADP, resulting in an inactive kinase. During infection by V.
parahaemolyticus, VopA/P efficiently blocks proliferative pathways while
allowing activation of survival pathways. Our understanding of this
infectious process is in its infancy; thus the activities of the other secreted
effectors will need to be uncovered to shed light on the importance of
VopA/P inhibition. Note that this type of serine/threonine acetylation has not
yet been observed as an endogenous protein modification in eukaryotes.
Figure 4. Multilevel regulation of small G proteins by bacterial effectors. Small G proteins cycle on
and off membranes, exchange guanine nucleotides (GDP to GTP) for activation, hydrolyze GTP for
inactivation, and stimulate downstream signaling pathways. Bacterial pathogens have evolved toxins
and effector proteins to usurp nearly every aspect of small G-protein function, using molecular mimicry
as well as novel stimulatory and inhibitory mechanisms.
7 CONCLUSION
Virulence factors produced by pathogens have evolved to efficiently
manipulate host signaling pathways (Table 1). Mechanisms range from
constitutive activation of a pathway, to irreversible inactivation of a critical
signaling molecule, to subversion of a whole signaling system to favor the
invading pathogen. A major challenge in the future is to determine the
enzymatic activities and host substrates for the bacterial and viral virulence
factors that show no obvious homology to eukaryotic proteins. Another, even
more complex challenge is to understand how these factors work together to
orchestrate a successful infection. Temporal and spatial considerations are
extremely important for regulating a host cell during infection. Likewise,
within the pathogen, determining the regulatory mechanisms that control the
activation patterns and spatial dynamics of virulence factors will help reveal
how microbial pathogens coopt signal transduction systems during infection.
Finally, the use of model organisms to complement studies in mammalian
cells will provide valuable insights into the physiological roles of bacterial
effector proteins. Such information is essential to gain a system-level view of
the infectious disease process and to ultimately design therapeutics that target
host–pathogen interactions. Inevitably, by discovering the mechanisms of
pathogenic effectors, we have gained a greater understanding into critical
steps in eukaryotic signaling. Although a great deal has been learned, given
the number and diversity of the yet-to-be-studied bacterial and viral
pathogens, much more is left to be discovered.
REFERENCES
*Reference is in this book.
Alto NM, Shao F, Lazar CS, Brost RL, Chua G, Mattoo S, McMahon SA, Ghosh P, Hughes TR, Boone
C, et al. 2006. Identification of a bacterial type III effector family with G protein mimicry functions.
Cell 124: 133–145.
Andor A, Trulzsch K, Essler M, Roggenkamp A, Wiedemann A, Heesemann J, Aepfelbacher M. 2001.
YopE of Yersinia, a GAP for Rho GTPases, selectively modulates Rac-dependent actin structures in
endothelial cells. Cell Microbiol 3: 301–310.
Black DS, Bliska JB. 2000. The RhoGAP activity of the Yersinia pseudotuberculosis cytotoxin YopE is
required for antiphagocytic function and virulence. Mol Microbiol 37: 515–527.
Broberg CA, Zhang L, Gonzalez H, Laskowski-Arce MA, Orth K. 2010. A Vibrio effector protein is an
inositol phosphatase and disrupts host cell membrane integrity. Science 329: 1660–1662.
Broberg CA, Calder TJ, Orth K. 2011. Vibrio parahaemolyticus cell biology and pathogenicity
determinants. Microbes Infect 13: 992–1001.
Brombacher E, Urwyler S, Ragaz C, Weber SS, Kami K, Overduin M, Hilbi H. 2009. Rab1 guanine
nucleotide exchange factor SidM is a major phosphatidylinositol 4-phosphate-binding effector
protein of Legionella pneumophila. J Biol Chem 284: 4846–4856.
Brown MS, Segal A, Stadtman ER. 1971. Modulation of glutamine synthetase adenylylation and
deadenylylation is mediated by metabolic transformation of the P II -regulatory protein. Proc Natl
Acad Sci 68: 2949–2953.
Buchwald G, Friebel A, Galan JE, Hardt WD, Wittinghofer A, Scheffzek K. 2002. Structural basis for
the reversible activation of a Rho protein by the bacterial toxin SopE. EMBO J 21: 3286–3295.
Campellone KG, Welch MD. 2010. A nucleator arms race: Cellular control of actin assembly. Nat Rev
11: 237–251.
Campellone KG, Robbins D, Leong JM. 2004. EspFU is a translocated EHEC effector that interacts
with Tir and N-WASP and promotes Nck-independent actin assembly. Dev Cell 7: 217–228.
Chardin P, Boquet P, Madaule P, Popoff MR, Rubin EJ, Gill DM. 1989. The mammalian G protein
rhoC is ADP-ribosylated by Clostridium botulinum exoenzyme C3 and affects actin microfilaments
in Vero cells. EMBO J 8: 1087–1092.
Charras G, Paluch E. 2008. Blebs lead the way: How to migrate without lamellipodia. Nat Rev 9: 730–
736.
Cheng HC, Skehan BM, Campellone KG, Leong JM, Rosen MK. 2008. Structural mechanism of
WASP activation by the enterohaemorrhagic E. coli effector EspF(U). Nature 454: 1009–1013.
Coburn J, Gill DM. 1991. ADP-ribosylation of p21ras and related proteins by Pseudomonas
aeruginosa exoenzyme S. Infect Immun 59: 4259–4262.
Coburn J, Wyatt RT, Iglewski BH, Gill DM. 1989. Several GTP-binding proteins, including p21c–H-
ras, are preferred substrates of Pseudomonas aeruginosa exoenzyme S. J Biol Chem 264: 9004–
9008.
Collier RJ. 2009. Membrane translocation by anthrax toxin. Mol Aspects Med 30: 413–422.
Collins CA, Brown EJ. 2010. Cytosol as battleground: Ubiquitin as a weapon for both host and
pathogen. Trends Cell Biol 20: 205–213.
Coscoy L, Ganem D. 2000. Kaposi’s sarcoma-associated herpesvirus encodes two proteins that block
cell surface display of MHC class I chains by enhancing their endocytosis. Proc Natl Acad Sci 97:
8051–8056.
Cui J, Yao Q, Li S, Ding X, Lu Q, Mao H, Liu L, Zheng N, Chen S, Shao F. 2010. Glutamine
deamidation and dysfunction of ubiquitin/NEDD8 induced by a bacterial effector family. Science
329: 1215–1218.
* Devreotes P, Horwitz AR. 2014. Signaling networks that regulate cell migration. Cold Spring Harb
Perspect Biol doi: 10.1101/cshperspect.a005959.
Dobner T, Horikoshi N, Rubenwolf S, Shenk T. 1996. Blockage by adenovirus E4orf6 of
transcriptional activation by the p53 tumor suppressor. Science 272: 1470–1473.
Egile C, Loisel TP, Laurent V, Li R, Pantaloni D, Sansonetti PJ, Carlier MF. 1999. Activation of the
CDC42 effector N-WASP by the Shigella flexneri IcsA protein promotes actin nucleation by
Arp2/3 complex and bacterial actin-based motility. J Cell Biol 146: 1319–1332.
Engel P, Goepfert A, Stanger FV, Harms A, Schmidt A, Schirmer T, Dehio C. 2012. Adenylylation
control by intra- or intermolecular active-site obstruction in Fic proteins. Nature 482: 107–110.
Etienne-Manneville S, Hall A. 2002. Rho GTPases in cell biology. Nature 420: 629–635.
Everett RD. 2000. ICP0, a regulator of herpes simplex virus during lytic and latent infection. Bioessays
22: 761–770.
Finn RD, Mistry J, Tate J, Coggill P, Heger A, Pollington JE, Gavin OL, Gunasekaran P, Ceric G,
Forslund K, et al. 2009. The Pfam protein families database. Nucleic Acids Res 38: D211–D222.
Flatau G, Lemichez E, Gauthier M, Chardin P, Paris S, Fiorentini C, Boquet P. 1997. Toxin-induced
activation of the G protein p21 Rho by deamidation of glutamine. Nature 387: 729–733.
Fu Y, Galan JE. 1999. A Salmonella protein antagonizes Rac-1 and Cdc42 to mediate host-cell
recovery after bacterial invasion. Nature 401: 293–297.
Garcia-Pino A, Christensen-Dalsgaard M, Wyns L, Yarmolinsky M, Magnuson RD, Gerdes K, Loris R.
2008. Doc of prophage P1 is inhibited by its antitoxin partner Phd through fold complementation. J
Biol Chem 283: 30821–30827.
Garmendia J, Phillips AD, Carlier MF, Chong Y, Schuller S, Marches O, Dahan S, Oswald E, Shaw
RK, Knutton S, et al. 2004. TccP is an enterohaemorrhagic Escherichia coli O157:H7 type III
effector protein that couples Tir to the actin-cytoskeleton. Cell Microbiol 6: 1167–1183.
Goehring UM, Schmidt G, Pederson KJ, Aktories K, Barbieri JT. 1999. The N-terminal domain of
Pseudomonas aeruginosa exoenzyme S is a GTPase-activating protein for Rho GTPases. J Biol
Chem 274: 36369–36372.
Gruenheid S, DeVinney R, Bladt F, Goosney D, Gelkop S, Gish GD, Pawson T, Finlay BB. 2001.
Enteropathogenic E. coli Tir binds Nck to initiate actin pedestal formation in host cells. Nat Cell
Biol 3: 856–859.
Guan KL, Dixon JE. 1990. Protein tyrosine phosphatase activity of an essential virulence determinant
in Yersinia. Science 249: 553–556.
Haas AL, Katzung DJ, Reback PM, Guarino LA. 1996. Functional characterization of the ubiquitin
variant encoded by the baculovirus Autographa californica. Biochemistry 35: 5385–5394.
Ham H, Sreelatha A, Orth K. 2011. Manipulation of host membranes by bacterial effectors. Nat Rev
Microbiol 9: 635–646.
Hao YH, Wang Y, Burdette D, Mukherjee S, Keitany G, Goldsmith E, Orth K. 2008. Structural
requirements for Yersinia YopJ inhibition of MAP kinase pathways. PLoS ONE 3: e1375.
Haque M, Ueda K, Nakano K, Hirata Y, Parravicini C, Corbellino M, Yamanishi K. 2001. Major
histocompatibility complex class I molecules are down-regulated at the cell surface by the K5
protein encoded by Kaposi’s sarcoma-associated herpesvirus/human herpesvirus-8. J Gen Virol 82:
1175–1180.
Hardt WD, Chen LM, Schuebel KE, Bustelo XR, Galan JE. 1998. S. typhimurium encodes an activator
of Rho GTPases that induces membrane ruffling and nuclear responses in host cells. Cell 93: 815–
826.
Hayes CS, Aoki SK, Low DA. 2010. Bacterial contact-dependent delivery systems. Annu Rev Genet
44: 71–90.
Henkel JS, Baldwin MR, Barbieri JT. 2010. Toxins from bacteria. EXS 100: 1–29.
Henriksson ML, Sundin C, Jansson AL, Forsberg A, Palmer RH, Hallberg B. 2002. Exoenzyme S
shows selective ADP-ribosylation and GTPase-activating protein (GAP) activities towards small
GTPases in vivo. Biochem J 367: 617–628.
Hernandez LD, Hueffer K, Wenk MR, Galan JE. 2004. Salmonella modulates vesicular traffic by
altering phosphoinositide metabolism. Science 304: 1805–1807.
Holt MR, Koffer A. 2001. Cell motility: Proline-rich proteins promote protrusions. Trends Cell Biol 11:
38–46.
Huang Z, Sutton SE, Wallenfang AJ, Orchard RC, Wu X, Feng Y, Chai J, Alto NM. 2009. Structural
insights into host GTPase isoform selection by a family of bacterial GEF mimics. Nat Struct Mol
Biol 16: 853–860.
Huibregtse JM, Scheffner M, Beaudenon S, Howley PM. 1995. A family of proteins structurally and
functionally related to the E6-AP ubiquitin-protein ligase. Proc Natl Acad Sci 92: 2563–2567.
Iriarte M, Cornelis GR. 1998. YopT, a new Yersinia Yop effector protein, affects the cytoskeleton of
host cells. Mol Microbiol 29: 915–929.
Ishido S, Wang C, Lee BS, Cohen GB, Jung JU. 2000. Downregulation of major histocompatibility
complex class I molecules by Kaposi’s sarcoma-associated herpesvirus K3 and K5 proteins. J Virol
74: 5300–5309.
Jiang P, Mayo AE, Ninfa AJ. 2007. Escherichia coli glutamine synthetase adenylyltransferase (ATase,
EC 2.7.7.49): Kinetic characterization of regulation by PII, PII-UMP, glutamine, and α-
ketoglutarate. Biochemistry 46: 4133–4146.
Jubelin G, Taieb F, Duda DM, Hsu Y, Samba-Louaka A, Nobe R, Penary M, Watrin C, Nougayrede JP,
Schulman BA, et al. 2010. Pathogenic bacteria target NEDD8-conjugated cullins to hijack host-cell
signaling pathways. PLoS Pathog 6: e1001128.
Juris SJ, Rudolph AE, Huddler D, Orth K, Dixon JE. 2000. A distinctive role for the Yersinia protein
kinase: Actin binding, kinase activation, and cytoskeleton disruption. Proc Natl Acad Sci 97: 9431–
9436.
Kenny B, DeVinney R, Stein M, Reinscheid DJ, Frey EA, Finlay BB. 1997. Enteropathogenic E. coli
(EPEC) transfers its receptor for intimate adherence into mammalian cells. Cell 91: 511–520.
Kinch LN, Yarbrough ML, Orth K, Grishin NV. 2009. Fido, a novel AMPylation domain common to
fic, doc, and AvrB. PLoS ONE 4: e5818.
Klink BU, Barden S, Heidler TV, Borchers C, Ladwein M, Stradal TE, Rottner K, Heinz DW. 2010.
Structure of Shigella IpgB2 in complex with human RhoA: Implications for the mechanism of
bacterial guanine nucleotide exchange factor mimicry. J Biol Chem 285: 17197–17208.
Kubori T, Galan JE. 2003. Temporal regulation of Salmonella virulence effector function by
proteasome-dependent protein degradation. Cell 115: 333–342.
* Lee MJ, Yaffe MB. 2014. Protein regulation in signal transduction. Cold Spring Harb Perspect Biol
doi: 10.1101/cshperspect.a005918.
Li H, Xu H, Zhou Y, Zhang J, Long C, Li S, Chen S, Zhou JM, Shao F. 2007. The phosphothreonine
lyase activity of a bacterial type III effector family. Science 315: 1000–1003.
* Lim K-H, Staudt LM. 2013. Toll-like receptor signaling. Cold Spring Harb Perspect Biol 5: a011247.
Liverman AD, Cheng HC, Trosky JE, Leung DW, Yarbrough ML, Burdette DL, Rosen MK, Orth K.
2007. Arp2/3-independent assembly of actin by Vibrio type III effector VopL. Proc Natl Acad Sci
104: 17117–17122.
Luong P, Kinch LN, Brautigam CA, Grishin NV, Tomchick DR, Orth K. 2010. Kinetic and structural
insights into the mechanism of AMPylation by VopS Fic domain. J Biol Chem 285: 20155–20163.
Miraglia AG, Travaglione S, Meschini S, Falzano L, Matarrese P, Quaranta MG, Viora M, Fiorentini
C, Fabbri A. 2007. Cytotoxic necrotizing factor 1 prevents apoptosis via the Akt/IκB kinase
pathway: Role of nuclear factor-κB and Bcl-2. Mol Biol Cell 18: 2735–2744.
Mittal R, Peak-Chew SY, McMahon HT. 2006. Acetylation of MEK2 and IκB kinase (IKK) activation
loop residues by YopJ inhibits signaling. Proc Natl Acad Sci 103: 18574–18579.
Mohr C, Koch G, Just I, Aktories K. 1992. ADP-ribosylation by Clostridium botulinum C3 exoenzyme
increases steady-state GTPase activities of recombinant rhoA and rhoB proteins. FEBS Lett 297:
95–99.
Morikawa H, Kim M, Mimuro H, Punginelli C, Koyama T, Nagai S, Miyawaki A, Iwai K, Sasakawa C.
2010. The bacterial effector Cif interferes with SCF ubiquitin ligase function by inhibiting
deneddylation of Cullin1. Biochem Biophys Res Commun 401: 268–274.
* Morrison DK. 2012. MAP kinase pathways. Cold Spring Harb Perspect Biol 4: a011254.
Mukherjee S, Keitany G, Li Y, Wang Y, Ball HL, Goldsmith EJ, Orth K. 2006. Yersinia YopJ
acetylates and inhibits kinase activation by blocking phosphorylation. Science 312: 1211–1214.
Mukherjee S, Hao YH, Orth K. 2007. A newly discovered post-translational modification—The
acetylation of serine and threonine residues. Trends Biochem Sci 32: 210–216.
Mukherjee S, Liu X, Arasaki K, McDonough J, Galan JE, Roy CR. 2011. Modulation of Rab GTPase
function by a protein phosphocholine transferase. Nature 477: 103–106.
Mukhopadhyay D, Riezman H. 2007. Proteasome-independent functions of ubiquitin in endocytosis
and signaling. Science 315: 201–205.
Muller MP, Peters H, Blumer J, Blankenfeldt W, Goody RS, Itzen A. 2010. The Legionella effector
protein DrrA AMPylates the membrane traffic regulator Rab1b. Science 329: 946–949.
Murata T, Delprato A, Ingmundson A, Toomre DK, Lambright DG, Roy CR. 2006. The Legionella
pneumophila effector protein DrrA is a Rab1 guanine nucleotide-exchange factor. Nat Cell Biol 8:
971–977.
Namgoong S, Boczkowska M, Glista MJ, Winkelman JD, Rebowski G, Kovar DR, Dominguez R.
2011. Mechanism of actin filament nucleation by Vibrio VopL and implications for tandem W
domain nucleation. Nat Struct Mol Biol 18: 1060–1067.
Navarro L, Koller A, Nordfelth R, Wolf-Watz H, Taylor S, Dixon JE. 2007. Identification of a
molecular target for the Yersinia protein kinase A. Mol Cell 26: 465–477.
Neunuebel MR, Chen Y, Gaspar AH, Backlund PS Jr, Yergey A, Machner MP. 2011. De-AMPylation
of the small GTPase Rab1 by the pathogen Legionella pneumophila. Science 333: 453–456.
* Newton K, Dixit VM. 2012. Signaling in innate immunity and inflammation. Cold Spring Harb
Perspect Biol 4: a006049.
Niebuhr K, Jouihri N, Allaoui A, Gounon P, Sansonetti PJ, Parsot C. 2000. IpgD, a protein secreted by
the type III secretion machinery of Shigella flexneri, is chaperoned by IpgE and implicated in entry
focus formation. Mol Microbiol 38: 8–19.
Niebuhr K, Giuriato S, Pedron T, Philpott DJ, Gaits F, Sable J, Sheetz MP, Parsot C, Sansonetti PJ,
Payrastre B. 2002. Conversion of PtdIns(4,5)P(2) into PtdIns(5)P by the S. flexneri effector IpgD
reorganizes host cell morphology. EMBO J 21: 5069–5078.
Norris FA, Wilson MP, Wallis TS, Galyov EE, Majerus PW. 1998. SopB, a protein required for
virulence of Salmonella dublin, is an inositol phosphate phosphatase. Proc Natl Acad Sci 95:
14057–14059.
Offermanns S, Toombs CF, Hu YH, Simon MI. 1997. Defective platelet activation in Gα(q)-deficient
mice. Nature 389: 183–186.
Orth K, Palmer LE, Bao ZQ, Stewart S, Rudolph AE, Bliska JB, Dixon JE. 1999. Inhibition of the
mitogen-activated protein kinase kinase superfamily by a Yersinia effector. Science 285: 1920–
1923.
Orth K, Xu Z, Mudgett MB, Bao ZQ, Palmer LE, Bliska JB, Mangel WF, Staskawicz B, Dixon JE.
2000. Disruption of signaling by Yersinia effector YopJ, a ubiquitin-like protein protease. Science
290: 1594–1597.
Pickart CM. 2004. Back to the future with ubiquitin. Cell 116: 181–190.
Prehna G, Ivanov MI, Bliska JB, Stebbins CE. 2006. Yersinia virulence depends on mimicry of host
Rho-family nucleotide dissociation inhibitors. Cell 126: 869–880.
Querido E, Blanchette P, Yan Q, Kamura T, Morrison M, Boivin D, Kaelin WG, Conaway RC,
Conaway JW, Branton PE. 2001a. Degradation of p53 by adenovirus E4orf6 and E1B55K proteins
occurs via a novel mechanism involving a Cullin-containing complex. Genes Dev 15: 3104–3117.
Querido E, Morrison MR, Chu-Pham-Dang H, Thirlwell SW, Boivin D, Branton PE. 2001b.
Identification of three functions of the adenovirus e4orf6 protein that mediate p53 degradation by
the E4orf6–E1B55K complex. J Virol 75: 699–709.
Randow F, Lehner PJ. 2009. Viral avoidance and exploitation of the ubiquitin system. Nat Cell Biol 11:
527–534.
Rigden DJ. 2011. Identification and modelling of a PPM protein phosphatase fold in the Legionella
pneumophila deAMPylase SidD. FEBS Lett 585: 2749–2754.
Rohde JR, Breitkreutz A, Chenal A, Sansonetti PJ, Parsot C. 2007. Type III secretion effectors of the
IpaH family are E3 ubiquitin ligases. Cell Host Microbe 1: 77–83.
Sallee NA, Rivera GM, Dueber JE, Vasilescu D, Mullins RD, Mayer BJ, Lim WA. 2008. The pathogen
protein EspF(U) hijacks actin polymerization using mimicry and multivalency. Nature 454: 1005–
1008.
* Sassone-Corsi P. 2012. The cyclic AMP pathway. Cold Spring Harb Perspect Biol 4: a011148.
Scheffner M, Huibregtse JM, Vierstra RD, Howley PM. 1993. The HPV-16 E6 and E6-AP complex
functions as a ubiquitin-protein ligase in the ubiquitination of p53. Cell 75: 495–505.
Schmidt G, Sehr P, Wilm M, Selzer J, Mann M, Aktories K. 1997. Gln 63 of Rho is deamidated by
Escherichia coli cytotoxic necrotizing factor-1. Nature 387: 725–729.
Sehr P, Joseph G, Genth H, Just I, Pick E, Aktories K. 1998. Glucosylation and ADP ribosylation of
rho proteins: Effects on nucleotide binding, GTPase activity, and effector coupling. Biochemistry
37: 5296–5304.
Selyunin AS, Sutton SE, Weigele BA, Reddick LE, Orchard RC, Bresson SM, Tomchick DR, Alto
NM. 2011. The assembly of a GTPase-kinase signalling complex by a bacterial catalytic scaffold.
Nature 469: 107–111.
Shao F, Merritt PM, Bao Z, Innes RW, Dixon JE. 2002. A Yersinia effector and a Pseudomonas
avirulence protein define a family of cysteine proteases functioning in bacterial pathogenesis. Cell
109: 575–588.
Shao F, Vacratsis PO, Bao Z, Bowers KE, Fierke CA, Dixon JE. 2003. Biochemical characterization of
the Yersinia YopT protease: Cleavage site and recognition elements in Rho GTPases. Proc Natl
Acad Sci 100: 904–909.
Singer AU, Rohde JR, Lam R, Skarina T, Kagan O, Dileo R, Chirgadze NY, Cuff ME, Joachimiak A,
Tyers M, et al. 2008. Structure of the Shigella T3SS effector IpaH defines a new class of E3
ubiquitin ligases. Nat Struct Mol Biol 15: 1293–1301.
Stebbins CE, Galan JE. 2000. Modulation of host signaling by a bacterial mimic: Structure of the
Salmonella effector SptP bound to Rac1. Mol Cell 6: 1449–1460.
Stevenson PG, Efstathiou S, Doherty PC, Lehner PJ. 2000. Inhibition of MHC class I-restricted antigen
presentation by γ2-herpesviruses. Proc Natl Acad Sci 97: 8455–8460.
Takai Y, Sasaki T, Matozaki T. 2001. Small GTP-binding proteins. Physiol Rev 81: 153–208.
Tan Y, Luo ZQ. 2011. Legionella pneumophila SidD is a deAMPylase that modifies Rab1. Nature 475:
506–509.
Terebiznik MR, Vieira OV, Marcus SL, Slade A, Yip CM, Trimble WS, Meyer T, Finlay BB, Grinstein
S. 2002. Elimination of host cell PtdIns(4,5)P(2) by bacterial SigD promotes membrane fission
during invasion by Salmonella. Nat Cell Biol 4: 766–773.
Trosky JE, Mukherjee S, Burdette DL, Roberts M, McCarter L, Siegel RM, Orth K. 2004. Inhibition of
MAPK signaling pathways by VopA from Vibrio parahaemolyticus. J Biol Chem 279: 51953–
51957.
Trosky JE, Li Y, Mukherjee S, Keitany G, Ball H, Orth K. 2007. VopA inhibits ATP binding by
acetylating the catalytic loop of MAPK kinases. J Biol Chem 282: 34299–34305.
Upadhyay A, Wu HL, Williams C, Field T, Galyov EE, van den Elsen JM, Bagby S. 2008. The
guanine-nucleotide-exchange factor BopE from Burkholderia pseudomallei adopts a compact
version of the Salmonella SopE/SopE2 fold and undergoes a closed-to-open conformational change
upon interaction with Cdc42. Biochem J 411: 485–493.
Von Pawel-Rammingen U, Telepnev MV, Schmidt G, Aktories K, Wolf-Watz H, Rosqvist R. 2000.
GAP activity of the Yersinia YopE cytotoxin specifically targets the Rho pathway: A mechanism
for disruption of actin microfilament structure. Mol Microbiol 36: 737–748.
Welch MD, Rosenblatt J, Skoble J, Portnoy DA, Mitchison TJ. 1998. Interaction of human Arp2/3
complex and the Listeria monocytogenes ActA protein in actin filament nucleation. Science 281:
105–108.
Woolery AR, Luong P, Broberg CA, Orth K. 2010. AMPylation: Something old is new again. Front
Microbiol 1: 113.
Worby CA, Mattoo S, Kruger RP, Corbeil LB, Koller A, Mendez JC, Zekarias B, Lazar C, Dixon JE.
2009. The Fic domain: Regulation of cell signaling by adenylylation. Mol Cell 34: 93–103.
Yarbrough ML, Li Y, Kinch LN, Grishin NV, Ball HL, Orth K. 2009. AMPylation of Rho GTPases by
Vibrio VopS disrupts effector binding and downstream signaling. Science 323: 269–272.
Yu Y, Wang SE, Hayward GS. 2005. The KSHV immediate-early transcription factor RTA encodes
ubiquitin E3 ligase activity that targets IRF7 for proteosome-mediated degradation. Immunity 22:
59–70.
Yu B, Cheng HC, Brautigam CA, Tomchick DR, Rosen MK. 2011. Mechanism of actin filament
nucleation by the bacterial effector VopL. Nat Struct Mol Biol 18: 1068–1074.
Zhu Y, Li H, Hu L, Wang J, Zhou Y, Pang Z, Liu L, Shao F. 2008. Structure of a Shigella effector
reveals a new class of ubiquitin ligases. Nat Struct Mol Biol 15: 1302–1308.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a006114
CHAPTER 21
Correspondence: joan_brugge@hms.harvard.edu
SUMMARY
Outline
1 Introduction
2 Mutations as the cause of cancer
3 Dysregulation of cellular processes by oncogenic signaling
4 Cell proliferation
5 Cell survival
6 Cell metabolism
7 Cell polarity and migration
8 Cell fate and differentiation
9 Genomic instability
10 The tumor microenvironment
11 Concluding remarks
References
1 INTRODUCTION
The development of cancer involves successive genetic and epigenetic
alterations that allow cells to escape homeostatic controls that ordinarily
suppress inappropriate proliferation and inhibit the survival of aberrantly
proliferating cells outside their normal niches. Most cancers arise in epithelial
cells, manifesting as carcinomas in organs such as the lung, skin, breast, liver,
and pancreas. Sarcomas, in contrast, arise from mesenchymal tissues,
occurring in fibroblasts, myocytes, adipocytes, and osteoblasts. Nonepithelial
tumors can also develop in cells of the nervous system (e.g., gliomas,
neuroblastomas, and medulloblastomas) and hematopoietic tissues (leukemia
and lymphoma).
In solid tumors, these alterations typically promote progression from a
relatively benign group of proliferating cells (hyperplasias) to a mass of cells
with abnormal morphology, cytological appearance, and cellular
organization. After a tumor expands, the tumor core loses access to oxygen
and nutrients, often leading to the growth of new blood vessels
(angiogenesis), which restores access to nutrients and oxygen. Subsequently,
tumor cells can develop the ability to invade the tissue beyond their normal
boundaries, enter the circulation, and seed new tumors at other locations
(metastasis), the defining feature of malignancy (Fig. 1). This linear sequence
of events is clearly an oversimplification of complex cancer-associated events
that proceed in distinct ways in individual tumors and between tumor sites;
however, it provides a useful framework in which to highlight the critical role
of dysregulated signaling in processes associated with the initiation and
progression of cancer.
4 CELL PROLIFERATION
Excessive cell proliferation is a feature of most cancers. Limited availability
of growth factors or nutrients, contact inhibition, and other feedback
mechanisms ensure that the pathways that regulate proliferation (see Fig. 3)
are normally tightly controlled. As outlined above, however, mutations in
proto-oncogenes and tumor suppressors or inappropriate synthesis of
ligands/receptors can hyperactivate these pathways, leading to activation of
the cell cycle machinery. Note that signaling targets that represent critical
components of cell cycle control mechanism can also undergo genetic
alterations in cancer; for example, the genes encoding cyclin D, cyclin E, and
CDK4 are amplified in certain cancers and the G1 restriction point inhibitor
pRB and p16 can be deleted or mutated as well.
Figure 3. Regulation of cell proliferation by the Ras-ERK and PI3K-Akt pathways.
ERK phosphorylates Bim and the NF-κB inhibitor IκBα (Ghoda et al.
1997), which targets them for degradation. In addition, RSK phosphorylates
the caspase-9 scaffolding protein APAF, which impedes the ability of
cytochrome c to nucleate apoptosome formation and activate the downstream
caspases that drive apoptosis (Kim et al. 2012).
6 CELL METABOLISM
Cell growth needs to be coordinated with metabolic processes involved in the
synthesis of macromolecules. Thus, growth factor pathways that regulate
both normal and tumor cells impinge on metabolic pathways to program cells
to meet the increased need for synthesis of macromolecules to produce new
daughter cells (Ch. 7 [Ward and Thompson 2012b]). Activation of oncogenes
and loss of tumor suppressors can directly regulate components of metabolic
pathways even in the absence of growth factors and, thereby, produce similar
metabolic alterations (Fig. 5).
Figure 5. Regulation of metabolism by Ras-ERK and PI3K-Akt signaling. 1DH*, mutated 1DH.
9 GENOMIC INSTABILITY
Genomic instability is a common characteristic of cancer cells. Aneuploidy
and large-scale DNA rearrangements are frequently observed, and many
cancers display elevated mutation rates. Ordinarily, a variety of cellular
enzymes repair DNA damage, and checkpoint signaling ensures that DNA
replication and cytokinesis are arrested in dividing cells until potentially
damaging errors are corrected. Alternatively, checkpoint signaling can induce
senescence or apoptosis so that affected cells do not pass on these errors.
Whether genomic instability is a cause or a consequence of cancer is still
debated, but it clearly reflects a failure of checkpoint signaling and/or DNA
repair mechanisms.
DNA damage signals are relayed by the kinases ATM, ATR, Chk1, and
Chk2, which stimulate p53, stall the cell cycle, and activate the DNA repair
machinery (p. 109 [Kopan 2012]; Ch. 6 [Rhind and Russell 2012]).
Downstream of p53, the CKI p21 is induced, and this can halt DNA
polymerase if DNA replication has already begun. If the damage cannot be
repaired and checkpoint signaling persists, p21 and p53 will induce cells to
senesce or undergo apoptosis (see above). Clearly, mutation or epigenetic
silencing of these tumor suppressors or upstream kinases can inactivate
checkpoint signaling, allowing DNA damage to persist and potentially fuel
cancer progression. Indeed, ATM and Chk2 mutations are seen in familial
leukemias and colon/breast cancers, respectively, and proteins involved in
DNA repair itself are also often mutated, for example, MMR enzymes and
BRCA1/2.
The mitotic checkpoint (also known as the spindle assembly checkpoint)
ensures that when a cell divides each daughter receives a full complement of
chromosomes. A complex containing the proteins Bub1, Bub3, and Mad1-3
monitors attachment of chromosomes to the mitotic spindle, relaying
checkpoint signals that block chromosome segregation and subsequent
cytokinesis. Once paired, sister chromatids are all attached to microtubules
emanating from opposite poles, the signal is switched off, and cells can move
from metaphase into anaphase and, ultimately, cytokinesis can proceed (Ch. 6
[Rhind and Russell 2012]). Inactivation of this checkpoint pathway has the
potential to lead to aneuploidy, and mutations in Mad1/2 and Bub1 have been
observed in cancer (Schvartzman et al. 2010).
Akt has been implicated in multiple aspects of DNA damage responses
and genome instability (Xu et al. 2012a). It can inhibit homologous
recombinational repair through direct phosphorylation of the checkpoint
proteins Chk1 and TopBP1 or indirectly through recruitment of resection
factors such as RPA, BRCA1, and Rad51 to sites of double-stranded breaks
(DSBs) in DNA. Akt is also activated by DSBs in a DNA-dependent protein-
kinase- or ATM/ATR-dependent manner and, in some contexts, can
contribute to radioresistance by stimulating DNA repair by nonhomologous
end joining. In addition, Akt also inhibits association of BRCA1 with DNA
damage foci. As discussed above, dysregulation of the PI3K-Akt pathway
suppresses apoptosis through many effectors, thus promoting survival of cells
with DNA damage. Because Ras-ERK signaling also inhibits apoptosis, it too
could promote survival of damaged cells. Hyperactivation of Ras-ERK
signaling has been shown to lead to genomic instability, although the
molecular mechanism is unclear (Saavedra et al. 1999). Akt therefore
modifies both the response to and repair of genotoxic damage in complex
ways that are likely to have important consequences for the therapy of tumors
showing deregulation of the PI3K-Akt pathway.
The tumor suppressor PTEN can also regulate chromosome stability,
independently of its 3′-phosphatase activity. PTEN regulates the expression
of the DNA repair protein RAD51, and loss of PTEN causes extensive
centromere breakage and chromosomal translocations (Toda et al. 1993; Liu
et al. 2008; p. 109 [Kopan 2012]).
Myc overexpression can induce genomic instability. In mammalian cells
and Drosophila, overexpression of Myc increases the frequency of
chromosomal rearrangements (Prochownik and Li 2007; Greer et al. 2013).
Multiple mechanisms have been associated with such genomic
rearrangements, including ROS-induced DSBs, suppression of checkpoints
that prevent replication of damaged DNA, and telomere clustering.
10 THE TUMOR MICROENVIRONMENT
So far, we have primarily considered how signaling within cancer cells
themselves is dysregulated in cancer. However, cancer progression (at least in
solid tumors) also depends on the ECM, blood vessels, immune cells, and
noncancerous cells such as fibroblasts in the tumor microenvironment, all of
which communicate with cancer cells by subverted signaling mechanisms
(Fig. 6). Many of the changes in the tumor microenvironment during cancer
progression mimic changes that occur during wound healing and/or
developmental processes. As tumors evolve, the complexity of their
“ecosystem” increases; reciprocal paracrine and juxtacrine interactions
between populations of neoplastic cells as well as tumor cells and
nonneoplastic cells within the microenvironment control cellular signaling
pathways in both positive and negative fashions. Dissecting the roles of
individual signaling pathways in these ecosystems is complex because it is
difficult to distinguish cell-autonomous and non-cell-autonomous activities.
Figure 6. Cancer signaling networks. The figure illustrates the wide variety of intra- and intercellular
signals affected in cancer, focusing on Ras-ERK and PI3K-Akt signaling. It is by no means
comprehensive; many more pathways are involved and there are other stromal cells involved in
paracrine signaling. Oncoproteins are indicated with yellow highlighting; tumor suppressors are
indicated with dashed outlines. Arrows do not necessarily indicate direct interactions in this figure.
10.2 Angiogenesis
Like all tissues, tumors require a blood supply. They acquire this by inducing
proliferation and assembly of endothelial cells to form new blood vessels
(angiogenesis), co-opting pathways that usually function in wound healing.
Central to angiogenesis are signals such as VEGF, PDGF, FGFs, interleukin
(IL) 8, and angiopoietin. The PI3K-Akt pathway regulates the induction of
angiogenesis as well as vessel integrity (Karar and Maity 2011). Synthesis
and secretion of VEGF by cancer cells is induced by HIF1. As mentioned
above, HIF1 levels are increased by PI3K-Akt signaling, and hyperactivation
of this pathway thus plays an important role in angiogenesis. HIF1 activity
can be controlled by the von Hippel–Lindau (VHL) protein, a subunit of an
E3 ligase that promotes its ubiquitin-dependent degradation under normoxic
conditions when it is proline hydroxylated by proline hydroxylases (see Ch. 7
[Ward and Thompson 2012b]), or through translational control. VHL
functions as a tumor suppressor and inactivating VHL mutations occur in a
variety of cancers. The PI3K-Akt pathway also modulates the production of
other angiogenic factors, such as nitric oxide and angiopoietins. Constitutive
endothelial activation of Akt1 has been shown to induce the formation of
structurally abnormal blood vessels.
Following its secretion, VEGF is sequestered in the ECM and cannot
exert its effects on endothelial cells until it is released by MMPs such as
MMP9. These are produced by monocytes and macrophages in the tumor
microenvironment (see below), which underscores the importance of immune
cells in angiogenesis and existence of the wider signaling network that
involves cancer cells, immune cells, and endothelial cells.
Another factor that must be overcome for angiogenesis to occur is
inhibitory signals such as thrombospondin 1 (Tsp1). Tsp1 released by various
cells normally keeps angiogenesis in check by inducing synthesis of FasL,
which causes endothelial cells to undergo apoptosis (see Ch. 19 [Green and
Llambi 2014]). Tsp1 is induced by p53, but repressed by Ras, Src, and Myc.
It thus represents another control point in angiogenesis that could be activated
by dysregulation of the Ras-ERK pathway. Moreover, the gene that encodes
Tsp1 is hypermethylated in some cancers.
10.3 Inflammation
Inflammatory cells such as macrophages and neutrophils constitute the first
defense against pathogens, but are also involved in tissue remodeling and
repair. They are recruited by chemokines secreted by tumor and stromal cells
to almost all tumors and secrete various molecules that promote cancer cell
proliferation, survival, and migration. In many respects, the contribution of
inflammatory cells to tumor progression, like that of angiogenesis,
recapitulates their role in wound healing, which also involves these
processes.
Signaling via the transcription factor NF-κB (see Ch. 15 [Newton and
Dixit 2012]) is important in both cancer cells and tumor-associated
inflammatory cells because it can promote cell survival and proliferation and
stimulates production of cytokines such as TNF. Oncogenic mutations
affecting NF-κB or upstream regulators such as MALT1 and Bcl10 occur in
some lymphoid malignancies; however, in most cancers, NF-κB activity is
simply increased by cytokine signaling. For example, in colon cancer, TNF
produced by macrophages increases NF-κB activity in intestinal epithelial
cells, which promotes cell survival; meanwhile, other cytokines such as IL6
and IL11 increase phospho-STAT3 levels (see p. 117 [Harrison 2012]),
which promote cell proliferation. A similar phenomenon occurs in
hepatocytes in hepatocellular carcinoma, the most common form of liver
cancer, and prostate cancer (Karin 2009). NF-κB activation also leads to
production of more TNF and synthesis of prostaglandin E2, which further
fuels cell proliferation and loss of cell polarity.
In addition to cytokines, inflammatory cells secrete growth factors such as
EGF and FGF. These are obviously important regulators of Ras-ERK and
PI3K-Akt signaling in cancer cells and will, therefore, dysregulate control of
cell proliferation, cell death, metabolism, and cell migration, as discussed
above. They also lead to production of colony-stimulating factor 1 (CSF1), a
key reciprocal signal that stimulates macrophages, causing them to produce
more EGF. Immune cells also produce VEGF and MMPs, which promotes
angiogenesis, ECM remodeling, and release of other bioactive molecules (see
above). Importantly, all these factors participate in paracrine loops involving
immune and cancer cells that sustain chronic inflammation and promote
tumor growth and progression.
11 CONCLUDING REMARKS
Cancer cells show a number of defining characteristics. Underlying these is a
dysregulation of cellular signal transduction induced by the genetic and
epigenetic changes that drive cancer. This affects not only the cancer cells
themselves, but the wider signaling network that encompasses other cells, the
ECM, blood vessels, and the immune system. Indeed, metastatic cancer can
be considered a systemic disease that affects signaling throughout the
affected individual, and systemic effects are ultimately what kill patients in
cancer.
Pharmacologic and antibody-based inhibitors that target signaling proteins
mutated in tumors or proteins downstream from these have had significant
impact as cancer treatments. For example, inhibitors of the nonreceptor
tyrosine kinase Abl and RTK ErbB2 dramatically reduce patient mortality in
chronic myelogenous leukemia and breast cancer. Other inhibitors, such as
those that target B-Raf, EGFR, and the kinase ALK induce remarkable
reductions of tumor volume and extend survival in patients with melanoma
and nonsmall-cell lung carcinomas; however, the rate of recurrence is high
because of the development of drug resistance (Gainor and Shaw 2013;
Giroux 2013; Holohan et al. 2013; Lito et al. 2013).
The complexity of the cancer signaling network (see Fig. 6) presents a
huge challenge for efforts to develop such anticancer drugs because of the
redundancy of pathways that control cell proliferation and survival, cross talk
between pathways, and feedback inhibition mechanisms that cause pathway
reactivation. The fact that pathways such as Ras-ERK and Akt-PI3K
signaling control so many characteristics of cancer cells, and that components
of these pathways, or upstream receptors, are so commonly mutated in a
variety of cancers gives reason to be optimistic that approaches based on
targeting them will be successful. The efficacy of therapies that target these
pathways is, however, limited by multiple factors. For example, rewiring of
signaling pathways is associated with adaptive responses to inhibition of
driver mutations, and this is commonly because of either loss of feedback
inhibition or induction of stress pathways (Pratilas and Solit, 2010; Rodrik-
Outmezguine et al. 2011; Lito et al. 2013). Moreover, factors from the tumor
microenvironment may stimulate alternative pathways that maintain cell
viability despite inhibition of the targeted pathways (Castells et al. 2012;
Muranen et al. 2012). Alternatively, there can be selection for rare tumor
cells that contain drug-resistant variants of the targeted protein or mutations
in other pathways that circumvent the dependency on the targeted pathway,
and epigenetic or stochastic changes in the state of tumor cells can also
activate intrinsic resistance pathways (Holohan et al. 2013; Holzel et al.
2013).
Further complicating matters is the degree of intratumoral genetic
heterogeneity. Recent evidence emerging from sequencing of single cells and
multiple regions of tumors from individual patients has revealed this is far
greater than previously imagined (Navin et al. 2011; Ruiz et al. 2011; Ding et
al. 2012; Gerlinger et al. 2012; Xu et al. 2012b; Bashashati et al. 2013) In one
study of kidney tumors, only ∼45% of mutations were detected in all tumor
regions. This heterogeneity also contributes significantly to intratumoral
variation in sensitivity to drugs targeting mutated signaling proteins mutated
in cancer, and means that single biopsies may not be sufficient to customize
patient treatment.
Overcoming these challenges will require a deeper understanding of the
nature of resistance mechanisms and how different cellular signaling
programs mediate resistant states in heterogeneous populations of tumor
cells. Combination therapies that target these should increase the efficacy of
targeted therapies. This is a significant challenge, but one that we feel is not
insurmountable.
REFERENCES
*Reference is in this book.
Anastasiou D, Poulogiannis G, Asara JM, Boxer MB, Jiang JK, Shen M, Bellinger G, Sasaki AT,
Locasale JW, Auld DS, et al. 2011. Inhibition of pyruvate kinase M2 by reactive oxygen species
contributes to cellular antioxidant responses. Science 334: 1278–1283.
Bakan I, Laplante M. 2012. Connecting mTORC1 signaling to SREBP-1 activation. Curr Opin Lipidol
23: 226–234.
Barthel A, Okino ST, Liao J, Nakatani K, Li J, Whitlock JP Jr., Roth RA. 1999. Regulation of GLUT1
gene transcription by the serine/threonine kinase Akt1. J Biol Chem 274: 20281–20286.
Bashashati A, Ha G, Tone A, Ding J, Prentice LM, Roth A, Rosner J, Shumansky K, Kalloger S, Senz
J, et al. 2013. Distinct evolutionary trajectories of primary high-grade serous ovarian cancers
revealed through spatial mutational profiling. J Pathol 231: 21–34.
Ben-Sahra I, Howell JJ, Asara JM, Manning BD. 2013. Stimulation of de novo pyrimidine synthesis by
growth signaling through mTOR and S6K1. Science 339: 1323–1328.
Berwick DC, Hers I, Heesom KJ, Moule SK, Tavare JM. 2002. The identification of ATP-citrate lyase
as a protein kinase B (Akt) substrate in primary adipocytes. J Biol Chem 277: 33895–33900.
Birchmeier W, Birchmeier C. 1995. Epithelial-mesenchymal transitions in development and tumor
progression. EXS 74: 1–15.
Burgering BM, Medema RH. 2003. Decisions on life and death: FOXO Forkhead transcription factors
are in command when PKB/Akt is off duty. J Leukoc Biol 73: 689–701.
Cagnol S, Chambard JC. 2010. ERK and cell death: Mechanisms of ERK-induced cell death—
Apoptosis, autophagy and senescence. FEBS J 277: 2–21.
Cain RJ, Ridley AJ. 2009. Phosphoinositide 3-kinases in cell migration. Biol Cell 101: 13–29.
Castells M, Thibault B, Delord JP, Couderc B. 2012. Implication of tumor microenvironment in
chemoresistance: Tumor-associated stromal cells protect tumor cells from cell death. Int J Mol Sci
13: 9545–9571.
Castoria G, Lombardi M, Barone MV, Bilancio A, Di Domenico M, De Falco A, Varricchio L, Bottero
D, Nanayakkara M, Migliaccio A, et al. 2004. Rapid signalling pathway activation by androgens in
epithelial and stromal cells. Steroids 69: 517–522.
Chin YR, Toker A. 2011. Akt isoform-specific signaling in breast cancer: Uncovering an anti-
migratory role for palladin. Cell Adh Migr 5: 211–214.
Christofk HR, Vander Heiden MG, Wu N, Asara JM, Cantley LC. 2008. Pyruvate kinase M2 is a
phosphotyrosine-binding protein. Nature 452: 181–186.
Cichowski K, Jacks T. 2001. NF1 tumor suppressor gene function: Narrowing the GAP. Cell 104: 593–
604.
Collak FK, Yagiz K, Luthringer DJ, Erkaya B, Cinar B. 2012. Threonine-120 phosphorylation
regulated by phosphoinositide-3-kinase/Akt and mammalian target of rapamycin pathway signaling
limits the antitumor activity of mammalian sterile 20-like kinase 1. J Biol Chem 287: 23698–23709.
Dang CV. 2013. MYC, metabolism, cell growth, and tumorigenesis. Cold Spring Harb Perspect Med
3: a014217.
Delhommeau F, Dupont S, Della Valle V, James C, Trannoy S, Masse A, Kosmider O, Le Couedic JP,
Robert F, Alberdi A, et al. 2009. Mutation in TET2 in myeloid cancers. N Engl J Med 360: 2289–
2301.
* Devreotes P, Horwitz AR. 2014. Signaling networks that regulate cell migration. Cold Spring Harb
Perspect Biol doi: 10.1101/cshperspect.a005959.
Diehl JA, Zindy F, Sherr CJ. 1997. Inhibition of cyclin D1 phosphorylation on threonine-286 prevents
its rapid degradation via the ubiquitin-proteasome pathway. Genes Dev 11: 957–972.
Ding L, Ley TJ, Larson DE, Miller CA, Koboldt DC, Welch JS, Ritchey JK, Young MA, Lamprecht T,
McLellan MD, et al. 2012. Clonal evolution in relapsed acute myeloid leukaemia revealed by
whole-genome sequencing. Nature 481: 506–510.
Doble BW, Woodgett JR. 2007. Role of glycogen synthase kinase-3 in cell fate and epithelial-
mesenchymal transitions. Cells Tissues Organs 185: 73–84.
Dow LE, Elsum IA, King CL, Kinross KM, Richardson HE, Humbert PO. 2008. Loss of human
Scribble cooperates with H-Ras to promote cell invasion through deregulation of MAPK signalling.
Oncogene 27: 5988–6001.
* Duronio RJ, Xiong Y. 2013. Signaling pathways that control cell proliferation. Cold Spring Harb
Perspect Biol 5: a008904.
Edinger AL, Thompson CB. 2002. Akt maintains cell size and survival by increasing mTOR-dependent
nutrient uptake. Mol Biol Cell 13: 2276–2288.
Elsum IA, Martin C, Humbert PO. 2013. Scribble regulates an EMT polarity pathway through
modulation of MAPK-ERK signaling to mediate junction formation. J Cell Sci 126: 3990–3999.
Engelman JA, Luo J, Cantley LC. 2006. The evolution of phosphatidylinositol 3-kinases as regulators
of growth and metabolism. Nat Rev Genet 7: 606–619.
Fang D, Hawke D, Zheng Y, Xia Y, Meisenhelder J, Nika H, Mills GB, Kobayashi R, Hunter T, Lu Z.
2007. Phosphorylation of β-catenin by AKT promotes β-catenin transcriptional activity. J Biol
Chem 282: 11221–11229.
Gainor JF, Shaw AT. 2013. Emerging paradigms in the development of resistance to tyrosine kinase
inhibitors in lung cancer. J Clin Oncol 31: 3987–3996.
Gao X, Wang H, Yang JJ, Liu X, Liu ZR. 2012. Pyruvate kinase M2 regulates gene transcription by
acting as a protein kinase. Mol Cell 45: 598–609.
Gerlinger M, Rowan AJ, Horswell S, Larkin J, Endesfelder D, Gronroos E, Martinez P, Matthews N,
Stewart A, Tarpey P, et al. 2012. Intratumor heterogeneity and branched evolution revealed by
multiregion sequencing. N Engl J Med 366: 883–892.
Ghoda L, Lin X, Greene WC. 1997. The 90-kDa ribosomal S6 kinase (pp90rsk) phosphorylates the N-
terminal regulatory domain of IκBα and stimulates its degradation in vitro. J Biol Chem 272:
21281–21288.
Giroux S. 2013. Overcoming acquired resistance to kinase inhibition: The cases of EGFR, ALK and
BRAF. Bioorg Med Chem Lett 23: 394–401.
Gorgoulis VG, Halazonetis TD. 2010. Oncogene-induced senescence: The bright and dark side of the
response. Curr Opin Cell Biol 22: 816–827.
* Green DR, Llambi F. 2014. Cell death signaling. Cold Spring Harb Perspect Biol doi:
10.1101/cshperspect.a006080.
Greer C, Lee M, Westerhof M, Milholland B, Spokony R, Vijg J, Secombe J. 2013. Myc-dependent
genome instability and lifespan in Drosophila. PLoS ONE 8: e74641.
Guo D, Bell EH, Mischel P, Chakravarti A. 2013. Targeting SREBP-1-driven lipid metabolism to treat
cancer. Curr Pharm Des. doi: 10.2174/13816128113199990486.
Hamede RK, McCallum H, Jones M. 2013. Biting injuries and transmission of Tasmanian devil facial
tumour disease. J Anim Ecol 82: 182–190.
Hanahan D, Weinberg RA. 2000. The hallmarks of cancer. Cell 100: 57–70.
Haq S, Michael A, Andreucci M, Bhattacharya K, Dotto P, Walters B, Woodgett J, Kilter H, Force T.
2003. Stabilization of β-catenin by a Wnt-independent mechanism regulates cardiomyocyte growth.
Proc Natl Acad Sci 100: 4610–4615.
* Hardie DG. 2012. Organismal carbohydrate and lipid homeostasis. Cold Spring Harb Perspect Biol
4: a006031.
* Harrison DA. 2012. The Jak/STAT pathway. Cold Spring Harb Perspect Biol 4: a011205.
* Harvey KF, Hariharan IK. 2012. The Hippo pathway. Cold Spring Harb Perspect Biol 4: a011288.
Harvey KF, Zhang X, Thomas DM. 2013. The Hippo pathway and human cancer. Nat Rev Cancer 13:
246–257.
* Hemmings BA, Restuccia DF. 2012. PI3K-PKB/Akt pathway. Cold Spring Harb Perspect Biol 4:
a011189.
Hitosugi T, Kang S, Vander Heiden MG, Chung TW, Elf S, Lythgoe K, Dong S, Lonial S, Wang X,
Chen GZ, et al. 2009. Tyrosine phosphorylation inhibits PKM2 to promote the Warburg effect and
tumor growth. Sci Signal 2: ra73.
Holohan C, Van Schaeybroeck S, Longley DB, Johnston PG. 2013. Cancer drug resistance: An
evolving paradigm. Nat Rev Cancer 13: 714–726.
Holzel M, Bovier A, Tuting T. 2013. Plasticity of tumour and immune cells: A source of heterogeneity
and a cause for therapy resistance? Nat Rev Cancer 13: 365–376.
Hu W, Zhang C, Wu R, Sun Y, Levine A, Feng Z. 2010. Glutaminase 2, a novel p53 target gene
regulating energy metabolism and antioxidant function. Proc Natl Acad Sci 107: 7455–7460.
Hussey GS, Chaudhury A, Dawson AE, Lindner DJ, Knudsen CR, Wilce MC, Merrick WC, Howe PH.
2011. Identification of an mRNP complex regulating tumorigenesis at the translational elongation
step. Mol Cell 41: 419–431.
Iden S, van Riel WE, Schafer R, Song JY, Hirose T, Ohno S, Collard JG. 2012. Tumor type-dependent
function of the par3 polarity protein in skin tumorigenesis. Cancer Cell 22: 389–403.
* Ingham PW. 2012. Hedgehog signaling. Cold Spring Harb Perspect Biol 4: a011221.
Israelsen WJ, Dayton TL, Davidson SM, Fiske BP, Hosios AM, Bellinger G, Li J, Yu Y, Sasaki M,
Horner JW, et al. 2013. PKM2 isoform-specific deletion reveals a differential requirement for
pyruvate kinase in tumor cells. Cell 155: 397–409.
Jensen PJ, Gunter LB, Carayannopoulos MO. 2010. Akt2 modulates glucose availability and
downstream apoptotic pathways during development. J Biol Chem 285: 17673–17680.
Jeon TI, Osborne TF. 2012. SREBPs: Metabolic integrators in physiology and metabolism. Trends
Endocrinol Metab 23: 65–72.
Kaelin WG Jr, McKnight SL. 2013. Influence of metabolism on epigenetics and disease. Cell 153: 56–
69.
Karar J, Maity A. 2011. PI3K/AKT/mTOR pathway in angiogenesis. Front Mol Neurosci 4: 51.
Karin M. 2009. NF-κB as a critical link between inflammation and cancer. Cold Spring Harb Perspect
Biol 1: a000141.
Keely PJ. 2011. Mechanisms by which the extracellular matrix and integrin signaling act to regulate the
switch between tumor suppression and tumor promotion. J Mammary Gland Biol Neoplasia 16:
205–219.
Kim JW, Zeller KI, Wang Y, Jegga AG, Aronow BJ, O’Donnell KA, Dang CV. 2004. Evaluation of
myc E-box phylogenetic footprints in glycolytic genes by chromatin immunoprecipitation assays.
Mol Cell Biol 24: 5923–5936.
Kim D, Shu S, Coppola MD, Kaneko S, Yuan ZQ, Cheng JQ. 2010. Regulation of proapoptotic
mammalian ste20-like kinase MST2 by the IGF1-Akt pathway. PLoS ONE 5: e9616.
Kim J, Parrish AB, Kurokawa M, Matsuura K, Freel CD, Andersen JL, Johnson CE, Kornbluth S.
2012. Rsk-mediated phosphorylation and 14-3-3ε binding of Apaf-1 suppresses cytochrome c-
induced apoptosis. EMBO J 31: 1279–1292.
* Kopan R. 2012. Notch signaling. Cold Spring Harb Perspect Biol 4: a011213.
Korkaya H, Paulson A, Charafe-Jauffret E, Ginestier C, Brown M, Dutcher J, Clouthier SG, Wicha
MS. 2009. Regulation of mammary stem/progenitor cells by PTEN/Akt/β-catenin signaling. PLoS
Biol 7: e1000121.
Kubatka P, Zihlavnikova K, Kajo K, Pec M, Stollarova N, Bojkova B, Kassayova M, Orendas P. 2011.
Antineoplastic effects of simvastatin in experimental breast cancer. Klin Onkol 24: 41–45.
* Laplante M, Sabatini DM. 2012. mTOR signaling. Cold Spring Harb Perspect Biol 4: a011593.
Larue L, Bellacosa A. 2005. Epithelial-mesenchymal transition in development and cancer: Role of
phosphatidylinositol 3′ kinase/AKT pathways. Oncogene 24: 7443–7454.
Ley TJ, Ding L, Walter MJ, McLellan MD, Lamprecht T, Larson DE, Kandoth C, Payton JE, Baty J,
Welch J, et al. 2010. DNMT3A mutations in acute myeloid leukemia. N Engl J Med 363: 2424–
2433.
Li S, Shen D, Shao J, Crowder R, Liu W, Prat A, He X, Liu S, Hoog J, Lu C, et al. 2013. Endocrine-
therapy-resistant ESR1 variants revealed by genomic characterization of breast-cancer-derived
xenografts. Cell Rep 4: 1116–1130.
Lin CY, Loven J, Rahl PB, Paranal RM, Burge CB, Bradner JE, Lee TI, Young RA. 2012.
Transcriptional amplification in tumor cells with elevated c-Myc. Cell 151: 56–67.
Lin JI, Poon CL, Harvey KF. 2013. The Hippo size control pathway-ever expanding. Sci Signal 6: pe4.
Lito P, Rosen N, Solit DB. 2013. Tumor adaptation and resistance to RAF inhibitors. Nat Med 19:
1401–1409.
Liu W, Zhou Y, Reske SN, Shen C. 2008. PTEN mutation: Many birds with one stone in
tumorigenesis. Anticancer Res 28: 3613–3619.
Lu P, Takai K, Weaver VM, Werb Z. 2011. Extracellular matrix degradation and remodeling in
development and disease. Cold Spring Harb Perspect Biol 3: a005058.
Luo W, Hu H, Chang R, Zhong J, Knabel M, O’Meally R, Cole RN, Pandey A, Semenza GL. 2011.
Pyruvate kinase M2 is a PHD3-stimulated coactivator for hypoxia-inducible factor 1. Cell 145:
732–744.
Ma L, Zhang G, Miao XB, Deng XB, Wu Y, Liu Y, Jin ZR, Li XQ, Liu QZ, Sun DX, et al. 2013.
Cancer stem-like cell properties are regulated by EGFR/AKT/β-catenin signaling and preferentially
inhibited by gefitinib in nasopharyngeal carcinoma. FEBS J 280: 2027–2041.
Mazzone M, Selfors LM, Albeck J, Overholtzer M, Sale S, Carroll DL, Pandya D, Lu Y, Mills GB,
Aster JC, et al. 2010. Dose-dependent induction of distinct phenotypic responses to Notch pathway
activation in mammary epithelial cells. Proc Natl Acad Sci 107: 5012–5017.
McCaffrey LM, Montalbano J, Mihai C, Macara IG. 2012. Loss of the Par3 polarity protein promotes
breast tumorigenesis and metastasis. Cancer Cell 22: 601–614.
McLaughlin SK, Olsen SN, Dake B, De Raedt T, Lim E, Bronson RT, Beroukhim R, Polyak K, Brown
M, Kuperwasser C, et al. 2013. The RasGAP gene, RASAL2, is a tumor and metastasis suppressor.
Cancer Cell 24: 365–378.
Miinea CP, Sano H, Kane S, Sano E, Fukuda M, Peranen J, Lane WS, Lienhard GE. 2005. AS160, the
Akt substrate regulating GLUT4 translocation, has a functional Rab GTPase-activating protein
domain. Biochem J 391: 87–93.
Miller DM, Thomas SD, Islam A, Muench D, Sedoris K. 2012. c-Myc and cancer metabolism. Clin
Cancer Res 18: 5546–5553.
Min J, Zaslavsky A, Fedele G, McLaughlin SK, Reczek EE, De Raedt T, Guney I, Strochlic DE,
Macconaill LE, Beroukhim R, et al. 2010. An oncogene-tumor suppressor cascade drives metastatic
prostate cancer by coordinately activating Ras and nuclear factor-κB. Nat Med 16: 286–294.
Morris MA, Dawson CW, Young LS. 2009. Role of the Epstein-Barr virus-encoded latent membrane
protein-1, LMP1, in the pathogenesis of nasopharyngeal carcinoma. Future Oncol 5: 811–825.
* Morrison DK. 2012. MAP kinase pathways. Cold Spring Harb Perspect Biol 4: a011254.
Munger K, Howley PM. 2002. Human papillomavirus immortalization and transformation functions.
Virus Res 89: 213–228.
Muranen T, Selfors LM, Worster DT, Iwanicki MP, Song L, Morales FC, Gao S, Mills GB, Brugge JS.
2012. Inhibition of PI3K/mTOR leads to adaptive resistance in matrix-attached cancer cells. Cancer
Cell 21: 227–239.
Murchison EP, Schulz-Trieglaff OB, Ning Z, Alexandrov LB, Bauer MJ, Fu B, Hims M, Ding Z,
Ivakhno S, Stewart C, et al. 2012. Genome sequencing and analysis of the Tasmanian devil and its
transmissible cancer. Cell 148: 780–791.
Murphy LO, Smith S, Chen RH, Fingar DC, Blenis J. 2002. Molecular interpretation of ERK signal
duration by immediate early gene products. Nat Cell Biol 4: 556–564.
Nagasaka K, Seiki T, Yamashita A, Massimi P, Subbaiah VK, Thomas M, Kranjec C, Kawana K,
Nakagawa S, Yano T, et al. 2013. A novel interaction between hScrib and PP1γ downregulates
ERK signaling and suppresses oncogene-induced cell transformation. PLoS ONE 8: e53752.
Nakashima A, Kawanishi I, Eguchi S, Yu EH, Eguchi S, Oshiro N, Yoshino K, Kikkawa U, Yonezawa
K. 2013. Association of CAD, a multifunctional protein involved in pyrimidine synthesis, with
mLST8, a component of the mTOR complexes. J Biomed Sci 20: 24.
Navin N, Kendall J, Troge J, Andrews P, Rodgers L, McIndoo J, Cook K, Stepansky A, Levy D,
Esposito D, et al. 2011. Tumour evolution inferred by single-cell sequencing. Nature 472: 90–94.
* Newton K, Dixit VM. 2012. Signaling in innate immunity and inflammation. Cold Spring Harb
Perspect Biol 4: a006049.
Nie Z, Hu G, Wei G, Cui K, Yamane A, Resch W, Wang R, Green DR, Tessarollo L, Casellas R, et al.
2012. c-Myc is a universal amplifier of expressed genes in lymphocytes and embryonic stem cells.
Cell 151: 68–79.
Novellasdemunt L, Tato I, Navarro-Sabate A, Ruiz-Meana M, Mendez-Lucas A, Perales JC, Garcia-
Dorado D, Ventura F, Bartrons R, Rosa JL. 2013. Akt-dependent activation of the heart 6-
phosphofructo-2-kinase/fructose-2,6-bisphosphatase (PFKFB2) isoenzyme by amino acids. J Biol
Chem 288: 10640–10651.
* Nusse R. 2012. Wnt signaling. Cold Spring Harb Perspect Biol 4: a011163.
Ogawara Y, Kishishita S, Obata T, Isazawa Y, Suzuki T, Tanaka K, Masuyama N, Gotoh Y. 2002. Akt
enhances Mdm2-mediated ubiquitination and degradation of p53. J Biol Chem 277: 21843–21850.
O’Neill E, Kolch W. 2005. Taming the Hippo: Raf-1 controls apoptosis by suppressing MST2/Hippo.
Cell Cycle 4: 365–367.
Plas DR, Thompson CB. 2005. Akt-dependent transformation: There is more to growth than just
surviving. Oncogene 24: 7435–7442.
Polakis P. 2001. More than one way to skin a catenin. Cell 105: 563–566.
Pratilas CA, Solit DB. 2010. Targeting the mitogen-activated protein kinase pathway: Physiological
feedback and drug response. Clin Cancer Res 16: 3329–3334.
Prochownik EV, Li Y. 2007. The ever expanding role for c-Myc in promoting genomic instability. Cell
Cycle 6: 1024–1029.
Raftopoulou M, Hall A. 2004. Cell migration: Rho GTPases lead the way. Dev Biol 265: 23–32.
Rahl PB, Lin CY, Seila AC, Flynn RA, McCuine S, Burge CB, Sharp PA, Young RA. 2010. c-Myc
regulates transcriptional pause release. Cell 141: 432–445.
Renoir JM, Marsaud V, Lazennec G. 2013. Estrogen receptor signaling as a target for novel breast
cancer therapeutics. Biochem Pharmacol 85: 449–465.
* Rhind N, Russell P. 2012. Signaling pathways that regulate cell division. Cold Spring Harb Perspect
Biol 4: a005942.
Richardson CJ, Schalm SS, Blenis J. 2004. PI3-kinase and TOR: PIKTORing cell growth. Semin Cell
Dev Biol 15: 147–159.
Roberts DJ, Tan-Sah VP, Smith JM, Miyamoto S. 2013. Akt phosphorylates HK-II at Thr-473 and
increases mitochondrial HK-II association to protect cardiomyocytes. J Biol Chem 288: 23798–
23806.
Robinson DR, Wu YM, Vats P, Su F, Lonigro RJ, Cao X, Kalyana-Sundaram S, Wang R, Ning Y,
Hodges L, et al. 2013. Activating ESR1 mutations in hormone-resistant metastatic breast cancer.
Nat Genet 45: 1446–1451.
Rodrik-Outmezguine VS, Chandarlapaty S, Pagano NC, Poulikakos PI, Scaltriti M, Moskatel E,
Baselga J, Guichard S, Rosen N. 2011. mTOR kinase inhibition causes feedback-dependent
biphasic regulation of AKT signaling. Cancer Discov 1: 248–259.
Rossig L, Jadidi AS, Urbich C, Badorff C, Zeiher AM, Dimmeler S. 2001. Akt-dependent
phosphorylation of pp21cip1 regulates PCNA binding and proliferation of endothelial cells. Mol
Cell Biol 21: 5644–5657.
Roux PP, Blenis J. 2004. ERK and p38 MAPK-activated protein kinases: A family of protein kinases
with diverse biological functions. Microbiol Mol Biol Rev 68: 320–344.
Ruiz C, Lenkiewicz E, Evers L, Holley T, Robeson A, Kiefer J, Demeure MJ, Hollingsworth MA, Shen
M, Prunkard D, et al. 2011. Advancing a clinically relevant perspective of the clonal nature of
cancer. Proc Natl Acad Sci 108: 12054–12059.
Saavedra HI, Fukasawa K, Conn CW, Stambrook PJ. 1999. MAPK mediates RAS-induced
chromosome instability. J Biol Chem 274: 38083–38090.
Sandoval J, Esteller M. 2012. Cancer epigenomics: Beyond genomics. Curr Opin Genet Dev 22: 50–
55.
Schvartzman JM, Sotillo R, Benezra R. 2010. Mitotic chromosomal instability and cancer: Mouse
modelling of the human disease. Nat Rev Cancer 10: 102–115.
Sears R, Nuckolls F, Haura E, Taya Y, Tamai K, Nevins JR. 2000. Multiple Ras-dependent
phosphorylation pathways regulate Myc protein stability. Genes Dev 14: 2501–2514.
Seeger C, Zoulim F, Mason W. 2013. Hepsdna viruses. In Fields virology, 3rd ed. (ed. Knipe DM,
Howley P). Lippincott Williams & Wilkins, Philadelphia.
* Sever R, Glass CK. 2013. Signaling by nuclear receptors. Cold Spring Harb Perspect Biol 5:
a016709.
Shaikh S, Collier DA, Sham PC, Ball D, Aitchison K, Vallada H, Smith I, Gill M, Kerwin RW. 1996.
Allelic association between a Ser-9-Gly polymorphism in the dopamine D3 receptor gene and
schizophrenia. Hum Genet 97: 714–719.
Shen HM, Tergaonkar V. 2009. NFκB signaling in carcinogenesis and as a potential molecular target
for cancer therapy. Apoptosis 14: 348–363.
Solimini NL, Luo J, Elledge SJ. 2007. Non-oncogene addiction and the stress phenotype of cancer
cells. Cell 130: 986–988.
Soloaga A, Thomson S, Wiggin GR, Rampersaud N, Dyson MH, Hazzalin CA, Mahadevan LC, Arthur
JS. 2003. MSK2 and MSK1 mediate the mitogen- and stress-induced phosphorylation of histone H3
and HMG-14. EMBO J 22: 2788–2797.
Sulzmaier FJ, Ramos JW. 2013. RSK isoforms in cancer cell invasion and metastasis. Cancer Res 73:
6099–6105.
Sutrias-Grau M, Arnosti DN. 2004. CtBP contributes quantitatively to Knirps repression activity in an
NAD binding-dependent manner. Mol Cell Biol 24: 5953–5966.
Suva ML, Riggi N, Bernstein BE. 2013. Epigenetic reprogramming in cancer. Science 339: 1567–1570.
Suzuki S, Tanaka T, Poyurovsky MV, Nagano H, Mayama T, Ohkubo S, Lokshin M, Hosokawa H,
Nakayama T, Suzuki Y, et al. 2010. Phosphate-activated glutaminase (GLS2), a p53-inducible
regulator of glutamine metabolism and reactive oxygen species. Proc Natl Acad Sci 107: 7461–
7466.
Thiery JP. 2002. Epithelial-mesenchymal transitions in tumour progression. Nat Rev Cancer 2: 442–
454.
Thomas C, Gustafsson JA. 2011. The different roles of ER subtypes in cancer biology and therapy. Nat
Rev Cancer 11: 597–608.
Toda M, Shirao T, Minoshima S, Shimizu N, Toya S, Uyemura K. 1993. Molecular cloning of cDNA
encoding human drebrin E and chromosomal mapping of its gene. Biochem Biophys Res Commun
196: 468–472.
Toy W, Shen Y, Won H, Green B, Sakr RA, Will M, Li Z, Gala K, Fanning S, King TA, et al. 2013.
ESR1 ligand-binding domain mutations in hormone-resistant breast cancer. Nat Genet 45: 1439–
1445.
Turner N, Grose R. 2010. Fibroblast growth factor signalling: From development to cancer. Nat Rev
Cancer 10: 116–129.
Turner SL, Blair-Zajdel ME, Bunning RA. 2009. ADAMs and ADAMTSs in cancer. Br J Biomed Sci
66: 117–128.
Vander Heiden MG, Cantley LC, Thompson CB. 2009. Understanding the Warburg effect: The
metabolic requirements of cell proliferation. Science 324: 1029–1033.
Vargas J, Feltes BC, Poloni Jde F, Lenz G, Bonatto D. 2012. Senescence: An endogenous anticancer
mechanism. Front Biosci (Landmark Ed) 17: 2616–2643.
Vicente-Manzanares M, Horwitz AR. 2011. Cell migration: An overview. Methods Mol Biol 769: 1–
24.
Vogelstein B, Papadopoulos N, Velculescu VE, Zhou S, Diaz LA Jr, Kinzler KW. 2013. Cancer
genome landscapes. Science 339: 1546–1558.
Ward PS, Thompson CB. 2012a. Metabolic reprogramming: A cancer hallmark even warburg did not
anticipate. Cancer Cell 21: 297–308.
* Ward PS, Thompson CB. 2012b. Signaling in control of cell growth and metabolism. Cold Spring
Harb Perspect Biol 4: a006783.
Weinberg R. 2013. The biology of cancer. Garland, New York.
Wieman HL, Wofford JA, Rathmell JC. 2007. Cytokine stimulation promotes glucose uptake via
phosphatidylinositol-3 kinase/Akt regulation of Glut1 activity and trafficking. Mol Biol Cell 18:
1437–1446.
Wu M, Pastor-Pareja JC, Xu T. 2010. Interaction between RasV12 and scribbled clones induces tumour
growth and invasion. Nature 463: 545–548.
Wu G, Broniscer A, McEachron TA, Lu C, Paugh BS, Becksfort J, Qu C, Ding L, Huether R, Parker
M, et al. 2012. Somatic histone H3 alterations in pediatric diffuse intrinsic pontine gliomas and
non-brainstem glioblastomas. Nat Genet 44: 251–253.
Xu N, Lao Y, Zhang Y, Gillespie DA. 2012a. Akt: A double-edged sword in cell proliferation and
genome stability. J Oncol 2012: 951724.
Xu X, Hou Y, Yin X, Bao L, Tang A, Song L, Li F, Tsang S, Wu K, Wu H, et al. 2012b. Single-cell
exome sequencing reveals single-nucleotide mutation characteristics of a kidney tumor. Cell 148:
886–895.
Xue B, Krishnamurthy K, Allred DC, Muthuswamy SK. 2013. Loss of Par3 promotes breast cancer
metastasis by compromising cell-cell cohesion. Nat Cell Biol 15: 189–200.
Yamada E, Okada S, Saito T, Ohshima K, Sato M, Tsuchiya T, Uehara Y, Shimizu H, Mori M. 2005.
Akt2 phosphorylates Synip to regulate docking and fusion of GLUT4-containing vesicles. J Cell
Biol 168: 921–928.
Ying H, Kimmelman AC, Lyssiotis CA, Hua S, Chu GC, Fletcher-Sananikone E, Locasale JW, Son J,
Zhang H, Coloff JL, et al. 2012. Oncogenic Kras maintains pancreatic tumors through regulation of
anabolic glucose metabolism. Cell 149: 656–670.
Yu FX, Guan KL. 2013. The Hippo pathway: Regulators and regulations. Genes Dev 27: 355–371.
Zhang X, Tang N, Hadden TJ, Rishi AK. 2011. Akt, FoxO and regulation of apoptosis. Biochim
Biophys Acta 1813: 1978–1986.
Cite this chapter as Cold Spring Harb Perspect Med doi: 10.1101/cshperspect.a006098
CHAPTER 22
Outlook
Correspondence: jthorner@berkeley.edu
SUMMARY
We have come a long way in the 55 years since Edmond Fischer and the
late Edwin Krebs discovered that the activity of glycogen phosphorylase
is regulated by reversible protein phosphorylation. As the contents of
this book attest, many of the other fundamental molecular mechanisms
that operate in biological signaling have since been characterized and the
vast web of interconnected pathways that make up the cellular signaling
network has been mapped in considerable detail. Nonetheless, it is
important to consider how fast this field is still moving and the issues at
the current boundaries of our understanding. One must also appreciate
what experimental strategies have allowed us to attain our present level
of knowledge. Therefore, we summarize here some key issues (both
conceptual and methodological), raise unresolved questions, discuss
potential pitfalls, and highlight areas in which our understanding is still
rudimentary. We hope these wide-ranging ruminations will be useful to
investigators who carry studies of signal transduction forward during the
rest of the 21st century.
Outline
1 Signaling in the atomic age
2 Seeing is believing
3 Signaling in the postgenomic era
4 Modularity in signaling
5 Posttranslational modifications and signaling
6 Integration of cell metabolism and signaling
7 Signal diversity
8 Prospectus
References
4 MODULARITY IN SIGNALING
One revelation derived from the explosion of sequence and structural
information is the extent to which signaling proteins appear to have arisen
during evolution by the shuffling and assembly of readily identifiable
modules. These stably folded domains, joined by flexible linkers, frequently
serve as recognition elements that mediate specific protein–protein
interactions that link them together in specific complexes. Understanding the
dynamics of the assembly of such complexes, how they direct signal
propagation, enhance signaling efficiency, and insulate pathways against
inadvertent stimulation continues to be an area of ongoing research.
Another important consequence of the modular nature of the proteins
involved in cellular regulation is that such an architecture allows the
constituent domains to evolve discrete and separable functions, which we are
just beginning to uncover and appreciate. For example, the p110β isoform of
the catalytic subunit of Class 1A phosphoinositide 3-kinase (PI3K) can
generate phosphatidylinositol 3,4,5-trisphosphate (PIP3) at the plasma
membrane via its carboxy-terminal kinase domain in response to growth-
factor-initiated signaling by receptor-tyrosine kinases or GPCR activation
(Vadas et al. 2011; Dbouk et al. 2012; see also p. 87 [Hemmings and
Restuccia 2012]). However, on growth factor withdrawal, p110β dissociates
from receptors, interacts via its so-called helical domain with the small
GTPase Rab5 and helps stabilize the GTP bound state of Rab5 (Dou et al.
2013). This interaction stimulates autophagy because increasing the amount
of Rab5-GTP that decorates internal membranes results in recruitment of the
class III PI3K hVps34, which generates the phosphatidylinositol 3-phosphate
necessary for assembly of preautophagosomes (Parzych and Klionsky 2013).
The mitogen-activated protein kinase (MAPK) kinase MEK1 provides
another example. It was thought to be simply a dedicated component of the
Ras-Raf-MEK-ERK signaling cascade (English and Cobb 2002; Roskoski
2012; see also p. 81 [Morrison 2012]); however, we now know that once
feedback phosphorylated on its amino-terminal extension by the kinase ERK,
MEK1 forms a ternary complex with a multidomain adaptor protein called
MAGI-1, which is necessary for membrane recruitment of the PIP3-specific
phosphatase PTEN; MEK1 thereby promotes down-regulation of the PIP3-
dependent protein kinase Akt (Zmajkovicova et al. 2013). A related issue is
that inherently disordered regions in some proteins adopt alternative
structures when associated with different interaction partners, leading to
different outcomes. The p53 transcription factor provides a particularly
dramatic example of this (Dunker et al. 2008; Joerger and Fersht 2008;
Freed-Pastor and Prives 2012).
The phenomenon whereby a protein has multiple distinct functions has
been dubbed “moonlighting” (Jeffery 2009). The potential for this evolving is
greatest in multidomain proteins, but not restricted to them. For example, one
of the splice variants of the muscle form of the glycolytic enzyme pyruvate
kinase, PKM2 (Hitosugi et al. 2009), appears to have other roles. On the one
hand, PKM2, when proline-hydroxylated by prolyl hydroxylase 3, seems to
associate with the transcription factor HIF1α and act as a coactivator that
promotes expression of HIF1α-dependent genes (Luo et al. 2011). On the
other hand, once tyrosine phosphorylated in response to growth factors,
PKM2 may undergo a switch in both oligomerization state (from a tetramer
to a dimer) and catalytic function (from its glycolytic role to a protein
kinase), and affect transcription by phosphorylating both histones (e.g., T11
on histone H3, which promotes acetylation at K9, a modification that
stimulates transcription) (Yang et al. 2012) and transcription factors (e.g.,
Y705 in STAT3, which promotes its dimerization and transactivator
function) (Gao et al. 2012).
Such instances of moonlighting in signaling proteins may help explain
how the intricacies of human biology are achieved with only 21,000 or so
protein-coding genes (just four times as many as a yeast cell). Thus,
investigators need to be alert to the possibility that moonlighting could
contribute in unanticipated ways to the biological complexity observed in a
signaling process, above and beyond alternative pre-mRNA splicing,
differential protein processing, and other mechanisms for generating protein
diversity that we already understand.
7 SIGNAL DIVERSITY
During evolution, mechanisms have arisen that allow diverse cell types to
sense and respond to various stimuli. In vision, light photons are absorbed by
the rhodopsins (chromophore-containing GPCRs) and converted into
intracellular chemical and then electrochemical changes in the rod and cone
cells in the retinas of our eyes. Hearing depends on conversion of sound
waves into intracellular signals via the opening and closing of stretch-
activated ion channels that are mechanically coupled to ciliary bundles in
specialized hair cells in our ears. In a similar way, touch depends on
conversion of mechanical pressure or thermal differences into conformational
changes in ion channels that open to elevate the level of intracellular calcium
ions in specialized nerve fibers in our skin. Of course, other stimuli to which
our senses respond are chemical in nature. Taste depends on conversion of
the binding of various soluble compounds into intracellular changes in the
gustatory-receptor-containing cells in the papillae on our tongues. Smell
depends on conversion of the binding of various volatile compounds into
intracellular signals in the olfactory-receptor-containing cells on the roof of
our nasal cavity. Skin irritation caused by the common nettle in our gardens
has a similar source; its trichomes inject chemicals normally made in animal
cells, such as the neurotransmitter serotonin and the immune mediator
histamine, thus aggravating our nerves and provoking inflammation.
Indeed, the variety of chemical signals generated by cells and to which
they are able to respond is rather staggering. Many of the classical endocrine
and pituitary hormones were discovered and chemically identified in the
1800s and early 1900s. For example, Frederick Banting announced the
isolation of insulin in 1921 and Fred Sanger determined its structure in 1953
(Nobel Prizes being awarded for both accomplishments). However, new
types of hormones that impact cell and organismal physiology and
development are still being discovered at a surprising rate, which requires
that their cognate receptors and downstream mediators be identified. For
example, adipose-derived leptin and its receptor (Isse et al. 1995; Tartaglia et
al. 1995), neuropeptides orexin-A/hypocretin-1 and andorexin-B/hypocretin-
2 and their receptors (de Lecea et al. 1998; Sakurai et al. 1998), and stomach-
derived ghrelin and its receptor (Howard et al. 1996; Kojima et al. 1999),
which have such critical roles in controlling the interrelated processes of
energy metabolism and obesity, wakefulness and appetite, and hunger and
growth, were not characterized until the late 1990s. Indeed, there are still
many GPCRs and nuclear receptors that are “orphans,” in the sense that their
physiological ligands have not yet been determined (Civelli 2012; Pearen and
Muscat 2012).
The constellation of known signaling molecules continues to expand in
new and unanticipated directions. For example, it has recently been
appreciated that different bacterial species, some of which are intracellular
pathogens, synthesize unusual cyclic dinucleotides that control their own
transcription, including cyclic-di-AMP, cyclic-di-GMP, and the mixed cyclic-
GMP-AMP, in all cases linked 3′-5′ (Kalia et al. 2013). The presence of such
compounds in mammalian cells is sensed by their binding to an endoplasmic
reticulum (ER)-localized protein called STING (Woodward et al. 2010).
Once activated, STING stimulates a protein kinase (TANK-binding kinase 1),
which, in turn, phosphorylates and activates a transcription factor, IRF3, that
regulates interferon production (Chin et al. 2013). However, STING is
activated much more potently and elicits a more efficacious interferon
response when foreign DNA enters cells. The difference is due to the fact that
cytosolic DNA binds to the regulatory domain of a mammalian enzyme,
cGAMP synthase (cGAS), thereby stimulating production of cyclic-GMP-
AMP in which the linkage is cyclic-[G(2′-5′)pA(3′-5′)p] (Kranzusch et al.
2013; Shaw and Liu 2014). This endogenously generated signal potently
activates diverse hSTING variants, whereas not all of them respond well to
bacterial cyclic-[G(3′-5′)pA(3′-5′)p] (Diner et al. 2013). Thus, mammalian
cells have evolved a mechanism to generate a novel signal that permits the
innate immune system to distinguish infiltration by a naked DNA from entry
of a bacterial invader. Such interkingdom signaling, mediating interplay
between viruses, bacteria, and mammalian cells (Lim et al. 2009; Marks et al.
2013; Pluznick et al. 2013), is clearly prevalent, has important consequences
for human health, and warrants continued exploration (see also Ch. 20 [Alto
and Orth 2012]).
How a signal is deployed or displayed can also have information content
that needs to be considered. Extracellular signaling ligands not only are
released from cells via the classical secretory pathway in an autocrine,
juxtacrine, or endocrine manner, but can also pass “through” cells via the
process of paracytophagy and the generation of so-called argosomes (Greco
et al. 2001). Such mechanisms are also used for the entry and cell-to-cell
passage of many prokaryotic intracellular pathogens (Portnoy 2012). Some
mammalian cell types can even engulf an entire other cell by a
macroendocytic process dubbed entosis (Overholtzer et al. 2007; Florey and
Overholtzer 2012). We now appreciate that cells can also generate
“exosomes” (small ∼100-nm-diameter vesicles) as a means for quantal export
of ligands and other classes of informational molecules, including miRNAs,
because, once released into extracellular fluids, they can be taken up by other
cells (Bang and Thum 2012; Briscoe and Thérond 2013; Choi et al. 2013).
Cells know their place, at least in part, by making contacts with adjacent
cells and components of the extracellular matrix. However, many cells types
erect a single specialized projection, the primary cilium (Garcia-Gonzalo and
Reiter 2012; Nozawa et al. 2013), which constrains to this one location
certain classes of signaling receptors (such as the receptors for the Hedgehog
family of ligands) (Wong and Reiter 2008; Ch. 10 [Perrimon et al. 2012]),
presumably to confer the capacity to respond only to a highly polarized or
localized signal source. In contrast, other cells extend ultrafine processes and
specialized filopodial extensions (also referred to as cytonemes) that can
mediate cell-to-cell contacts and long-range transport of signal molecules
over a distance of many cell lengths (Roy et al. 2011; Sanders et al. 2013).
Similarly, juxtaposed cells in many epithelial layers are connected by gap
junctions that act as portholes through which certain intracellular signals,
such as an increase in cytosolic calcium ion concentration, can be spread
from cell to cell (Goodenough and Paul 2009). Clearly, we still have much to
learn about the interplay between signaling molecules and all levels of
cellular organization.
8 PROSPECTUS
For the foreseeable future, signal transduction research remains a field
confronted with a still vast frontier. As the contents of this book and issues
raised here make clear, unraveling signal transduction processes is a
multidisciplinary enterprise. Ultimately, understanding signaling will require
an appreciation for, understanding of, and the means to usefully grasp the
seamless interconnectedness among all the bits of information gleaned by
what were formerly considered disparate branches of the biological,
chemical, and physical sciences. Fortunately, as we highlight in this book,
there are recurrent themes, general mechanisms, common strategies, and
ubiquitous reactions in cell signaling that allow the complexity to be parsed
out productively. Indeed, we can already discern general design principles
that are allowing us to reengineer cellular responses to stimuli different from
those evolved in nature (Pryciak 2009; Burrill et al. 2011; Blount et al. 2012;
Lim et al. 2013).
Nonetheless, many unexpected discoveries that provide novel insights and
fresh paradigms will continue to be made, and these will open up new
avenues for understanding cell signaling. Some recent examples highlight
this point. Remarkably, proteins of the previously uncharacterized Fam20C
family turn out to be secreted atypical protein kinases that phosphorylate
casein and other extracellular substrates that have important physiological
roles in bone mineralization (Tagliabracci et al. 2013; Xiao et al. 2013b).
This discovery raises interesting questions about how these enzymes acquire
ATP in the lumen of the Golgi and whether such enzymes are made and have
important roles in the nervous system, in which ATP is stored in synaptic
vesicles (Zimmermann 2008) and released extracellularly to stimulate
purinergic receptors (Khakh and North 2012). Likewise, it has been
appreciated for decades that blood platelets also store ATP and other
compounds in storage vesicles that are released extracellularly when platelets
are activated at a site of injury (Da Prada et al. 1971; Higashii et al. 1985).
Does this released ATP also act, in part, through Fam20C-like extracellular
protein kinases? Another example is that very high serum levels of high-
molecular-weight hyaluronan dramatically protect the naked mole rat against
cancer (Tian et al. 2013). Hyaluronan is an extracellular proteoglycan and can
bind to various cell surface receptors. How does it suppress malignant growth
and/or enhance immune surveillance of precancerous tissue? And, do the
same mechanisms operate in humans?
It is also clear that we have just began to scratch the surface of how
microRNAs (Martinez and Gregory 2010; Mendell and Olson 2012) and
other RNAs encoded in our genomes (Hancks and Kazazian 2012; Batista
and Chang 2013) coordinately influence the levels of gene products involved
in signaling and are themselves controlled by signaling processes. Likewise,
an area of traditional biochemistry that clearly intersects with signaling in
many ways is the function of various classes of proteases; yet, we understand
the roles of many of these enzymes only very superficially and know even
less about their physiologically relevant enzyme–substrate relationships. For
example, in 2012, a new, circulating, exercise-induced regulatory hormone,
dubbed irisin, was described that converts white fat into more thermogenic
beige fat (Boström et al. 2012). However, irisin is identical to the ectodomain
of a small (212-residue) cell surface protein, fibronectin type III domain-
containing protein 5 (FNDC5), which is anchored in the plasma membrane
by a single carboxy-terminal hydrophobic transmembrane segment. In fact,
protease-mediated shedding of the extracellular domains of transmembrane
signaling proteins as separate entities with distinct functions is a common
phenomenon (Horiuchi 2013). Exercise stimulates FNDC5 expression in
skeletal muscle and the cleavage and release of its uniquely structured amino-
terminal fibronectin-III-like ectodomain as irisin (Erickson 2013;
Schumacher et al. 2013); however, the protease responsible for this shedding,
and whether it too is under any sort of regulation, is unknown.
Thus, if we take the long view, studies of signal transduction are still in
exponential phase with many important discoveries to come. Moreover, we
anticipate continued development of evermore sophisticated experimental
tools—from improvements in automated deep sequencing to characterize the
global transcriptome (Malone and Oliver 2011), to new mass spectrometry
instrumentation to catalog the cellular metabolome (Rubakhin et al. 2013), to
further refinement of mathematical, statistical, and computational theories
and methods to assist with display, interpretation, and modeling of the
complex networks of relationships involved in intra- and intercellular
signaling (Janes and Lauffenburger 2013; see also Ch. 4 [Azeloglu and
Iyengar 2014]). Continued advances of this sort will allow us to address
questions at an ever greater level of detail and resolution, providing answers
at the molecular level to long-standing mechanistic questions about the
myriad processes that comprise cell signaling.
REFERENCES
*Reference is in this book.
Akerboom J, Chen TW, Wardill TJ, Tian L, Marvin JS, Mutlu S, Calderón NC, Esposti F, Borghuis
BG, Sun XR, et al. 2012. Optimization of a GCaMP calcium indicator for neural activity imaging. J
Neurosci 32: 13819–13840.
Allen JJ, Li M, Brinkworth CS, Paulson JL, Wang D, Hübner A, Chou WH, Davis RJ, Burlingame AL,
Messing RO, et al. 2007. A semi-synthetic epitope for kinase substrates. Nat Methods 4: 511–516.
* Alto NM, Orth K. 2012. Subversion of cell signaling by pathogens. Cold Spring Harb Perspect Biol
4: a006114.
Arsenault R, Griebel P, Napper S. 2011. Peptide arrays for kinome analysis: New opportunities and
remaining challenges. Proteomics 11: 4595–4609.
* Azeloglu EU, Iyengar R. 2014. Signaling networks: Information flow, computation, and decision
making. Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a005934.
Baker D. 2006. Prediction and design of macromolecular structures and interactions. Philos Trans R
Soc Lond B Biol Sci 361: 459–463.
Bandura DR, Baranov VI, Ornatsky OI, Antonov A, Kinach R, Lou X, Pavlov S, Vorobiev S, Dick JE,
Tanner SD. 2009. Mass cytometry: Technique for real time single cell multitarget immunoassay
based on inductively coupled plasma time-of-flight mass spectrometry. Anal Chem 81: 6813–6822.
Bang C, Thum T. 2012. Exosomes: New players in cell-cell communication. Int J Biochem Cell Biol
44: 2060–2064.
Bardwell L, Thorner J. 1996. A conserved motif at the amino termini of MEKs might mediate high-
affinity interaction with the cognate MAPKs. Trends Biochem Sci 21: 373–374.
Bargmann CI, Kaplan JM. 1998. Signal transduction in the Caenorhabditis elegans nervous system.
Annu Rev Neurosci 21: 279–308.
Bar-Peled L, Schweitzer LD, Zoncu R, Sabatini DM. 2012. Ragulator is a GEF for the rag GTPases
that signal amino acid levels to mTORC1. Cell 150: 1196–1208.
Batista PJ, Chang HY. 2013. Long noncoding RNAs: Cellular address codes in development and
disease. Cell 152: 1298–1307.
Baumeister W, Steven AC. 2000. Macromolecular electron microscopy in the era of structural
genomics. Trends Biochem Sci 25: 624–631.
Beauchamp KA, McGibbon R, Lin YS, Pande VS. 2012. Simple few-state models reveal hidden
complexity in protein folding. Proc Natl Acad Sci 109: 17807–17813.
Bendall SC, Simonds EF, Qiu P, Amir el-AD, Krutzik PO, Finck R, Bruggner RV, Melamed R, Trejo
A, Ornatsky OI, et al. 2011. Single-cell mass cytometry of differential immune and drug responses
across a human hematopoietic continuum. Science 332: 687–696.
Bensimon A, Heck AJ, Aebersold R. 2012. Mass spectrometry-based proteomics and network biology.
Annu Rev Biochem 81: 379–405.
Berndt N, Hamilton AD, Sebti SM. 2011. Targeting protein prenylation for cancer therapy. Nat Rev
Cancer 11: 775–791.
Bessman NJ, Lemmon MA. 2012. Finding the missing links in EGFR. Nat Struct Mol Biol 19: 1–3.
Bhattacharyya RP, Reményi A, Yeh BJ, Lim WA. 2006. Domains, motifs, and scaffolds: The role of
modular interactions in the evolution and wiring of cell signaling circuits. Annu Rev Biochem 75:
655–680.
Blanchetot C, Chagnon M, Dubé N, Hallé M, Tremblay ML. 2005. Substrate-trapping techniques in the
identification of cellular PTP targets. Methods 35: 44–53.
Blount BA, Weenink T, Ellis T. 2012. Construction of synthetic regulatory networks in yeast. FEBS
Lett 586: 2112–2121.
Bodenmiller B, Zunder ER, Finck R, Chen TJ, Savig ES, Bruggner RV, Simonds EF, Bendall SC,
Sachs K, Krutzik PO, et al. 2012. Multiplexed mass cytometry profiling of cellular states perturbed
by small-molecule regulators. Nat Biotechnol 30: 858–867.
Borrmann A, Milles S, Plass T, Dommerholt J, Verkade JM, Wiessler M, Schultz C, van Hest JC, van
Delft FL, Lemke EA. 2012. Genetic encoding of a bicyclo[6.1.0]nonyne-charged amino acid
enables fast cellular protein imaging by metal-free ligation. Chembiochem 13: 2094–2099.
Boström P, Wu J, Jedrychowski MP, Korde A, Ye L, Lo JC, Rasbach KA, Boström EA, Choi JH, Long
JZ, et al. 2012. A PGC1-α-dependent myokine that drives brown-fat-like development of white fat
and thermogenesis. Nature 481: 463–468.
Boubekeur S, Boute N, Pagesy P, Zilberfarb V, Christeff M, Issad T. 2011. A new highly efficient
substrate-trapping mutant of protein tyrosine phosphatase 1B (PTP1B) reveals full autoactivation of
the insulin receptor precursor. J Biol Chem 286: 19373–19380.
Bradshaw JM. 2010. The Src, Syk, and Tec family kinases: Distinct types of molecular switches. Cell
Signal 22: 1175–1184.
Briscoe J, Thérond PP. 2013. The mechanisms of Hedgehog signalling and its roles in development and
disease. Nat Rev Mol Cell Biol 14: 416–429.
Brooks BR, Brooks CLIII, Mackerell ADJ, Nilsson L, Petrella RJ, Roux B, Won Y, Archontis G,
Bartels C, Boresch S, et al. 2009. CHARMM: The biomolecular simulation program. J Comput
Chem 30: 1545–1614.
Burrill DR, Boyle PM, Silver PA. 2011. A new approach to an old problem: Synthetic biology tools for
human disease and metabolism. Cold Spring Harb Symp Quant Biol 76: 145–154.
Butko MT, Yang J, Geng Y, Kim HJ, Jeon NL, Shu X, Mackey MR, Ellisman MH, Tsien RY, Lin MZ.
2012. Fluorescent and photo-oxidizing TimeSTAMP tags track protein fates in light and electron
microscopy. Nat Neurosci 15: 1742–1751.
Carroll D. 2011. Genome engineering with zinc-finger nucleases. Genetics 188: 773–782.
Chakraborty AK, Das J. 2010. Pairing computation with experimentation: A powerful coupling for
understanding T cell signalling. Nat Rev Immunol 19: 59–71.
Chalfie M, Tu Y, Euskirchen G, Ward WW, Prasher DC. 1994. Green fluorescent protein as a marker
for gene expression. Science 263: 802–805.
Cheong R, Wang CJ, Levchenko A. 2009. Using a microfluidic device for high-content analysis of cell
signaling. Sci Signal 2: pl2.1–pl2.18.
Chi A, Huttenhower C, Geer LY, Coon JJ, Syka JE, Bai DL, Shabanowitz J, Burke DJ, Troyanskaya
OG, Hunt DF. 2007. Analysis of phosphorylation sites on proteins from Saccharomyces cerevisiae
by electron transfer dissociation (ETD) mass spectrometry. Proc Natl Acad Sci 104: 2193–2198.
Chin JW. 2014. Expanding and reprogramming the genetic code of cells and animals. Annu Rev
Biochem doi: 10.1146/annurev-biochem-060713-035737.
Chin KH, Tu ZL, Su YC, Yu YJ, Chen HC, Lo YC, Chen CP, Barber GN, Chuah ML, Liang ZX, et al.
2013. Novel c-di-GMP recognition modes of the mouse innate immune adaptor protein STING.
Acta Crystallogr D Biol Crystallogr 69: 352–366.
Chiu W, Baker ML, Almo SC. 2006. Structural biology of cellular machines. Trends Cell Biol 16: 144–
150.
Choi DS, Kim DK, Kim YK, Gho YS. 2013. Proteomics, transcriptomics and lipidomics of exosomes
and ectosomes. Proteomics 13: 1554–1571.
Christie JM, Gawthorne J, Young G, Fraser NJ, Roe AJ. 2012. LOV to BLUF: Flavoprotein
contributions to the optogenetic toolkit. Mol Plant 5: 533–544.
Civelli O. 2012. Orphan GPCRs and neuromodulation. Neuron 76: 12–21.
Close DM, Xu T, Sayler GS, Ripp S. 2011. In vivo bioluminescent imaging (BLI): Noninvasive
visualization and interrogation of biological processes in living animals. Sensors 11: 180–206.
Coelho M, Maghelli N, Tolic-Nørrelykke IM. 2013. Single-molecule imaging in vivo: The dancing
building blocks of the cell. Integr Biol 5: 748–758.
Cohen P. 1992. Signal integration at the level of protein kinases, protein phosphatases and their
substrates. Trends Biochem Sci 17: 408–413.
Cohen P. 2002. The origins of protein phosphorylation. Nat Cell Biol 4: E127–E130.
Cohen P, Alessi DR. 2013. Kinase drug discovery—What’s next in the field? ACS Chem Biol 8: 96–
104.
Cohen P, Knebel A. 2006. KESTREL: A powerful method for identifying the physiological substrates
of protein kinases. Biochem J 393: 1–6.
Cook JG, Bardwell L, Kron SJ, Thorner J. 1996. Two novel targets of the MAP kinase Kss1 are
negative regulators of invasive growth in the yeast Saccharomyces cerevisiae. Genes Dev 10:
2831–2848.
Copley RR, Doerks T, Letunic I, Bork P. 2002. Protein domain analysis in the era of complete
genomes. FEBS Lett 513: 129–134.
Corrêa IR Jr, Baker B, Zhang A, Sun L, Provost CR, Lukinavicius G, Reymond L, Johnsson K, Xu
MQ. 2013. Substrates for improved live-cell fluorescence labeling of SNAP-tag. Curr Pharm Des
19: 5414–5420.
Crabtree GR, Schreiber SL. 1996. Three-part inventions: Intracellular signaling and induced proximity.
Trends Biochem Sci 21: 418–422.
Creanga A, Glenn TD, Mann RK, Saunders AM, Talbot WS, Beachy PA. 2012. Scube/You activity
mediates release of dually lipid-modified Hedgehog signal in soluble form. Genes Dev 26: 1312–
1325.
Dang L, White DW, Gross S, Bennett BD, Bittinger MA, Driggers EM, Fantin VR, Jang HG, Jin S,
Keenan MC, et al. 2009. Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature
462: 739–744.
Da Prada M, Pletscher A, Tranzer JP. 1971. Storage of ATP and 5-hydroxytryptamine in blood
platelets of guinea-pigs. J Physiol 217: 679–688.
Dbouk HA, Vadas O, Williams RL, Backer JM. 2012. PI3Kβ downstream of GPCRs—Crucial partners
in oncogenesis. Oncotarget 3: 1485–1486.
Debets MF, van Hest JC, Rutjes FP. 2013. Bioorthogonal labelling of biomolecules: New functional
handles and ligation methods. Org Biomol Chem 11: 6439–6455.
de Lecea L, Kilduff TS, Peyron C, Gao X, Foye PE, Danielson PE, Fukuhara C, Battenberg EL,
Gautvik VT, Bartlett FS, et al. 1998. The hypocretins: Hypothalamus-specific peptides with
neuroexcitatory activity. Proc Natl Acad Sci 95: 322–327.
Dennis EA, Cao J, Hsu YH, Magrioti V, Kokotos G. 2011. Phospholipase A2 enzymes: Physical
structure, biological function, disease implication, chemical inhibition, and therapeutic intervention.
Chem Rev 111: 6130–6185.
Deshaies RJ, Emberley ED, Saha A. 2010. Control of cullin-ring ubiquitin ligase activity by nedd8.
Subcell Biochem 54: 41–56.
Dewey CN, Pachter L. 2006. Evolution at the nucleotide level: The problem of multiple whole-genome
alignment. Hum Mol Genet 15: R51–R56.
Diner EJ, Burdette DL, Wilson SC, Monroe KM, Kellenberger CA, Hyodo M, Hayakawa Y, Hammond
MC, Vance RE. 2013. The innate immune DNA sensor cGAS produces a noncanonical cyclic
dinucleotide that activates human STING. Cell Rep 3: 1355–1361.
Doolittle RF. 1995. The multiplicity of domains in proteins. Annu Rev Biochem 64: 287–314.
Dou Z, Pan JA, Dbouk HA, Ballou LM, DeLeon JL, Fan Y, Chen JS, Liang Z, Li G, Backer JM, et al.
2013. Class IA PI3K p110β subunit promotes autophagy through Rab5 small GTPase in response to
growth factor limitation. Mol Cell 50: 29–42.
Dror RO, Dirks RM, Grossman JP, Xu H, Shaw DE. 2012. Biomolecular simulation: A computational
microscope for molecular biology. Annu Rev Biophys 41: 429–452.
Duncan JS, Whittle MC, Nakamura K, Abell AN, Midland AA, Zawistowski JS, Johnson NL, Granger
DA, Jordan NV, Darr DB, et al. 2012. Dynamic reprogramming of the kinome in response to
targeted MEK inhibition in triple-negative breast cancer. Cell 149: 307–321.
Dunker AK, Silman I, Uversky VN, Sussman JL. 2008. Function and structure of inherently disordered
proteins. Curr Opin Struct Biol 18: 756–764.
* Duronio RJ, Xiong Y. 2013. Signaling pathways that control cell proliferation. Cold Spring Harb
Perspect Biol 5: a008904.
Elphick LM, Lee SE, Gouverneur V, Mann DJ. 2007. Using chemical genetics and ATP analogues to
dissect protein kinase function. ACS Chem Biol 2: 299–314.
Encell LP, Friedman Ohana R, Zimmerman K, Otto P, Vidugiris G, Wood MG, Los GV, McDougall
MG, Zimprich C, Karassina N, et al. 2012. Development of a dehalogenase-based protein fusion tag
capable of rapid, selective and covalent attachment to customizable ligands. Curr Chem Genomics
6: 55–71.
Endicott JA, Noble MA, Johnson LN. 2012. The structural basis for control of eukaryotic protein
kinases. Annu Rev Biochem 81: 587–613.
Endres NF, Das R, Smith AW, Arkhipov A, Kovacs E, Huang Y, Pelton JG, Shan Y, Shaw DE,
Wemmer DE, et al. 2013. Conformational coupling across the plasma membrane in activation of
the EGF receptor. Cell 152: 543–556.
Engholm-Keller K, Larsen MR. 2013. Technologies and challenges in large-scale phosphoproteomics.
Proteomics 13: 910–931.
English JM, Cobb MH. 2002. Pharmacological inhibitors of MAPK pathways. Trends Pharmacol Sci
23: 40–45.
Erickson HP. 2013. Irisin and FNDC5 in retrospect: An exercise hormone or a transmembrane
receptor? Adipocyte 2: 289–293.
Farley AR, Link AJ. 2009. Identification and quantification of protein posttranslational modifications.
Methods Enzymol 463: 725–763.
Farrar MA, Alberol-Ila J, Perlmutter RM. 1996. Activation of the Raf-1 kinase cascade by
coumermycin-induced dimerization. Nature 383: 178–181.
Feldman ME, Shokat KM. 2010. New inhibitors of the PI3K-Akt-mTOR pathway: Insights into mTOR
signaling from a new generation of Tor kinase domain inhibitors (TORKinibs). Curr Top Microbiol
Immunol 347: 241–262.
Fenno L, Yizhar O, Deisseroth K. 2011. The development and application of optogenetics. Annu Rev
Neurosci 34: 389–412.
Ferrell JE Jr. 2000. What do scaffold proteins really do? Sci STKE 2000: e1.
Filonov GS, Verkhusha VV. 2013. A near-infrared BiFC reporter for in vivo imaging of protein-protein
interactions. Chem Biol 20: 1078–1086.
Fischer EH. 2013. Cellular regulation by protein phosphorylation. Biochem Biophys Res Commun 430:
865–867.
Florey O, Overholtzer M. 2012. Autophagy proteins in macroendocytic engulfment. Trends Cell Biol
22: 374–380.
Frame S, Cohen P. 2001. GSK3 takes centre stage more than 20 years after its discovery. Biochem J
359: 1–16.
Frank J. 2009. Single-particle reconstruction of biological macromolecules in electron microscopy—30
years. Q Rev Biophys 42: 139–158.
Freed-Pastor WA, Prives C. 2012. Mutant p53: One name, many proteins. Genes Dev 26: 1268–1286.
Freeman RS, Hasbani DM, Lipscomb EA, Straub JA, Xie L. 2003. SM-20, EGL-9, and the EGLN
family of hypoxia-inducible factor prolyl hydroxylases. Mol Cells 16: 1–12.
Friedland GD, Kortemme T. 2010. Designing ensembles in conformational and sequence space to
characterize and engineer proteins. Curr Opin Struct Biol 20: 377–384.
Fukada M, Noda M. 2007. Yeast substrate-trapping system for isolating substrates of protein tyrosine
phosphatases. Methods Mol Biol 365: 371–382.
Gaj T, Gersbach CA, Barbas CFIII. 2013. ZFN, TALEN, and CRISPR/Cas-based methods for genome
engineering. Trends Biotechnol 31: 397–405.
Galperin MY, Koonin EV. 2012. Divergence and convergence in enzyme evolution. J Biol Chem 287:
21–28.
Gao X, Wang H, Yang JJ, Liu X, Liu ZR. 2012. Pyruvate kinase M2 regulates gene transcription by
acting as a protein kinase. Mol Cell 45: 598–609.
Garcia-Gonzalo FR, Reiter JF. 2012. Scoring a backstage pass: Mechanisms of ciliogenesis and ciliary
access. J Cell Biol 197: 697–709.
Gautier A, Juillerat A, Heinis C, Corrêa IRJ, Kindermann M, Beaufils F, Johnsson K. 2008. An
engineered protein tag for multiprotein labeling in living cells. Chem Biol 15: 128–136.
Gevaert K, Vandekerckhove J. 2009. Reverse-phase diagonal chromatography for phosphoproteome
research. Methods Mol Biol 527: 219–227.
Ghosh M, Tucker DE, Burchett SA, Leslie CC. 2006. Properties of the Group IV phospholipase A2
family. Prog Lipid Res 45: 487–510.
Goldsmith EJ, Akella R, Min X, Zhou T, Humphreys JM. 2007. Substrate and docking interactions in
serine/threonine protein kinases. Chem Rev 107: 5065–5081.
González-Vera JA. 2012. Probing the kinome in real time with fluorescent peptides. Chem Soc Rev 41:
1652–1664.
Goodenough DA, Paul DL. 2009. Gap junctions. Cold Spring Harb Perspect Biol 1: a002576.
Gorostiza P, Isacoff EY. 2008. Optical switches for remote and noninvasive control of cell signaling.
Science 322: 395–399.
Granier S, Kobilka B. 2012. A new era of GPCR structural and chemical biology. Nat Chem Biol 8:
670–673.
Greco V, Hannus M, Eaton S. 2001. Argosomes: A potential vehicle for the spread of morphogens
through epithelia. Cell 106: 633–645.
Greenleaf WJ, Woodside MT, Block SM. 2007. High-resolution, single-molecule measurements of
biomolecular motion. Annu Rev Biophys Biomol Struct 36: 171–190.
Groves JT, Kuriyan J. 2010. Molecular mechanisms in signal transduction at the membrane. Nat Struct
Mol Biol 17: 659–665.
Guerra C, Barbacid M. 2013. Genetically engineered mouse models of pancreatic adenocarcinoma. Mol
Oncol 7: 232–247.
Guerrero G, Isacoff EY. 2001. Genetically encoded optical sensors of neuronal activity and cellular
function. Curr Opin Neurobiol 11: 601–607.
Guibert S, Weber M. 2013. Functions of DNA methylation and hydroxymethylation in mammalian
development. Curr Top Dev Biol 104: 47–83.
Hancks DC, Kazazian HHJ. 2012. Active human retrotransposons: Variation and disease. Curr Opin
Genet Dev 22: 191–203.
* Hardie DG. 2012. Organismal carbohydrate and lipid homeostasis. Cold Spring Harb Perspect Biol
4: a006031.
Hart GW, Slawson C, Ramirez-Correa G, Lagerlof O. 2011. Cross talk between O-GlcNAcylation and
phosphorylation: Roles in signaling, transcription, and chronic disease. Annu Rev Biochem 80: 825–
858.
Harvey CD, Ehrhardt AG, Cellurale C, Zhong H, Yasuda R, Davis RJ, Svoboda K. 2008. A genetically
encoded fluorescent sensor of ERK activity. Proc Natl Acad Sci 105: 19264–19269.
* Hemmings BA, Restuccia DF. 2012. The PI3K-PKB/Akt pathway. Cold Spring Harb Perspect Biol
4: a011189.
Herner A, Nikic I, Kállay M, Lemke EA, Kele P. 2013. A new family of bioorthogonally applicable
fluorogenic labels. Org Biomol Chem 11: 3297–3306.
Hertz NT, Wang BT, Allen JJ, Zhang C, Dar AC, Burlingame AL, Shokat KM. 2010. Chemical genetic
approach for kinase-substrate mapping by covalent capture of thiophosphopeptides and analysis by
mass spectrometry. Curr Protoc Chem Biol 2: 15–36.
Higashii T, Isomoto A, Tyuma I, Kakishita E, Uomoto M, Nagai K. 1985. Quantitative and continuous
analysis of ATP release from blood platelets with firefly luciferase luminescence. Thromb Haemost
53: 65–69.
Hitosugi T, Kang S, Vander Heiden MG, Chung TW, Elf S, Lythgoe K, Dong S, Lonial S, Wang Z,
Chen GZ, et al. 2009. Tyrosine phosphorylation inhibits PKM2 to promote the Warburg effect and
tumor growth. Sci Signal 2: ra73.71–ra73.78.
Horiuchi K. 2013. A brief history of tumor necrosis factor α-converting enzyme: An overview of
ectodomain shedding. Keio J Med 62: 29–36.
Hou H, Yu H. 2010. Structural insights into histone lysine demethylation. Curr Opin Struct Biol 20:
739–748.
Hou F, Sun L, Zheng H, Skaug B, Jiang QX, Chen ZJ. 2011. MAVS forms functional prion-like
aggregates to activate and propagate antiviral innate immune response. Cell 146: 448–461.
Howard AD, Feighner SD, Cully DF, Arena JP, Liberator PA, Rosenblum CI, Hamelin M, Hreniuk
DL, Palyha OC, Anderson J, et al. 1996. A receptor in pituitary and hypothalamus that functions in
growth hormone release. Science 273: 974–977.
Huang B, Bates M, Zhuang X. 2009. Super-resolution fluorescence microscopy. Annu Rev Biochem 78:
993–1016.
Huang IK, Pei J, Grishin NV. 2013. Defining and predicting structurally conserved regions in protein
superfamilies. Bioinformatics 29: 175–181.
Hunter T. 2012. Why nature chose phosphate to modify proteins. Philos Trans R Soc Lond B Biol Sci
367: 2513–2516.
* Ingham PW. 2012. Hedgehog signaling. Cold Spring Harb Perspect Biol 4: a011221.
Isse N, Ogawa Y, Tamura N, Masuzaki H, Mori K, Okazaki T, Satoh N, Shigemoto M, Yoshimasa Y,
Nishi S, et al. 1995. Structural organization and chromosomal assignment of the human obese gene.
J Biol Chem 270: 27728–27733.
Janes KA, Lauffenburger DA. 2013. Models of signalling networks—What cell biologists can gain
from them and give to them. J Cell Sci 126: 1913–1921.
Jeffery CJ. 2009. Moonlighting proteins—An update. Mol Biosyst 5: 345–350.
Jiang Z, Rokhsar DS, Harland RM. 2009. Old can be new again: HAPPY whole genome sequencing,
mapping and assembly. Int J Biol Sci 5: 298–303.
Jin J, Pawson T. 2012. Modular evolution of phosphorylation-based signalling systems. Philos Trans R
Soc Lond B Biol Sci 367: 2540–2555.
Joerger AC, Fersht AR. 2008. Structural biology of the tumor suppressor p53. Annu Rev Biochem 77:
557–582.
Jura N, Zhang X, Endres NF, Seeliger MA, Schindler T, Kuriyan J. 2011. Catalytic control in the EGF
receptor and its connection to general kinase regulatory mechanisms. Mol Cell 42: 9–22.
Kalia D, Merey G, Nakayama S, Zheng Y, Zhou J, Luo Y, Guo M, Roembke BT, Sintim HO. 2013.
Nucleotide, c-di-GMP, c-di-AMP, cGMP, cAMP, (p)ppGpp signaling in bacteria and implications
in pathogenesis. Chem Soc Rev 42: 305–341.
Kang JY, Kawaguchi D, Coin I, Xiang Z, O’Leary DD, Slesinger PA, Wang L. 2013. In vivo
expression of a light-activatable potassium channel using unnatural amino acids. Neuron 80: 358–
370.
Katritch V, Cherezov V, Stevens RC. 2013. Structure-function of the G protein-coupled receptor
superfamily. Annu Rev Pharmacol Toxicol 53: 531–556.
Kerppola TK. 2009. Visualization of molecular interactions using bimolecular fluorescence
complementation analysis: Characteristics of protein fragment complementation. Chem Soc Rev 38:
2876–2886.
Khakh BS, North RA. 2012. Neuromodulation by extracellular ATP and P2X receptors in the CNS.
Neuron 76: 51–69.
Kiu H, Nicholson SE. 2012. Biology and significance of the JAK/STAT signalling pathways. Growth
Factors 30: 88–106.
Kleiman LB, Maiwald T, Conzelmann H, Lauffenburger DA, Sorger PK. 2011. Rapid phospho-
turnover by receptor tyrosine kinases impacts downstream signaling and drug binding. Mol Cell 43:
723–737.
Kliegman JI, Fiedler D, Ryan CJ, Xu YF, Su XY, Thomas D, Caccese MC, Cheng A, Shales M,
Rabinowitz JD, et al. 2013. Chemical genetics of rapamycin-insensitive TORC2 in S. cerevisiae.
Cell Rep 5: 1725–1736.
Knight ZA, Shokat KM. 2007. Chemical genetics: Where genetics and pharmacology meet. Cell 128:
425–430.
Knight JD, Pawson T, Gingras AC. 2013. Profiling the kinome: Current capabilities and future
challenges. J Proteomics 81: 43–55.
Knighton DR, Zheng JH, Ten Eyck LF, Ashford VA, Xuong NH, Taylor SS, Sowadski JM. 1991.
Crystal structure of the catalytic subunit of cyclic adenosine monophosphate-dependent protein
kinase. Science 253: 407–414.
Kõivomägi M, Valk E, Venta R, Iofik A, Lepiku M, Balog ER, Rubin SM, Morgan DO, Loog M. 2011.
Cascades of multisite phosphorylation control Sic1 destruction at the onset of S phase. Nature 480:
128–131.
Kojima M, Hosoda H, Date Y, Nakazato M, Matsuo H, Kangawa K. 1999. Ghrelin is a growth-
hormone-releasing acylated peptide from stomach. Nature 402: 656–660.
Kolb P, Ferreira RS, Irwin JJ, Shoichet BK. 2009. Docking and chemoinformatic screens for new
ligands and targets. Curr Opin Biotechnol 20: 429–436.
Kramer RH, Mourot A, Adesnik H. 2013. Optogenetic pharmacology for control of native neuronal
signaling proteins. Nat Neurosci 16: 816–823.
Kranzusch PJ, Lee AS, Berger JM, Doudna JA. 2013. Structure of human cGAS reveals a conserved
family of second-messenger enzymes in innate immunity. Cell Rep 3: 1362–1368.
Krishna RG, Wold F. 1993. Post-translational modification of proteins. Adv Enzymol Relat Areas Mol
Biol 67: 265–298.
Kumar J, Mayer ML. 2013. Functional insights from glutamate receptor ion channel structures. Annu
Rev Physiol 75: 313–337.
Kunkel MT, Toker A, Tsien RY, Newton AC. 2007. Calcium-dependent regulation of protein kinase D
revealed by a genetically encoded kinase activity reporter. J Biol Chem 282: 6733–6742.
Kuriyan J, Eisenberg D. 2007. The origin of protein interactions and allostery in colocalization. Nature
450: 983–990.
Lander GC, Saibil HR, Nogales E. 2012. Go hybrid: EM, crystallography, and beyond. Curr Opin
Struct Biol 22: 627–635.
Lane TJ, Shukla D, Beauchamp KA, Pande VS. 2013. To milliseconds and beyond: Challenges in the
simulation of protein folding. Curr Opin Struct Biol 23: 58–65.
Lassila JK, Zalatan JG, Herschlag D. 2011. Biological phosphoryl-transfer reactions: Understanding
mechanism and catalysis. Annu Rev Biochem 80: 669–702.
Lau SY, Procko E, Gaudet R. 2012. Distinct properties of Ca2+-calmodulin binding to N- and C-
terminal regulatory regions of the TRPV1 channel. J Gen Physiol 140: 541–555.
* Lee ML, Yaffe MB. 2014. Protein regulation in signal transduction. Cold Spring Harb Perspect Biol
doi: 10.1101/cshperspect.a005918.
Leitner A, Sturm M, Lindner W. 2011. Tools for analyzing the phosphoproteome and other
phosphorylated biomolecules. Anal Chim Acta 703: 19–30.
Leivar P, Quail PH. 2011. PIFs: Pivotal components in a cellular signaling hub. Trends Plant Sci 16:
19–28.
Levskaya A, Weiner OD, Lim WA, Voigt CA. 2009. Spatiotemporal control of cell signalling using a
light-switchable protein interaction. Nature 461: 997–1001.
Li M, Yu Y, Yang J. 2011. Structural biology of TRP channels. Adv Exp Med Biol 704: 1–23.
Lim JH, Kim HJ, Komatsu K, Ha U, Huang Y, Jono H, Kweon SM, Lee J, Xu X, Zhang GS, et al.
2009. Differential regulation of Streptococcus pneumoniae-induced human MUC5AC mucin
expression through distinct MAPK pathways. Am J Transl Res 1: 300–311.
Lim WA, Lee CM, Tang C. 2013. Design principles of regulatory networks: Searching for the
molecular algorithms of the cell. Mol Cell 49: 202–212.
Lima SQ, Miesenböck G. 2005. Remote control of behavior through genetically targeted
photostimulation of neurons. Cell 121: 141–152.
Lin H, Su X, He B. 2012. Protein lysine acylation and cysteine succination by intermediates of energy
metabolism. ACS Chem Biol 7: 947–960.
Lin R, Tao R, Gao X, Li T, Zhou X, Guan KL, Xiong Y, Lei QY. 2013. Acetylation stabilizes ATP-
citrate lyase to promote lipid biosynthesis and tumor growth. Mol Cell 51: 506–518.
Liu CC, Schultz PG. 2010. Adding new chemistries to the genetic code. Annu Rev Biochem 79: 413–
444.
Londoño Gentile T, Lu C, Lodato PM, Tse S, Olejniczak SH, Witze ES, Thompson CB, Wellen KE.
2013. DNMT1 is regulated by ATP-citrate lyase and maintains methylation patterns during
adipocyte differentiation. Mol Cell Biol 33: 3864–3878.
Loroch S, Dickhut C, Zahedi RP, Sickmann A. 2013. Phosphoproteomics—More than meets the eye.
Electrophoresis 34: 1483–1492.
Losman JA, Kaelin WG Jr. 2013. What a difference a hydroxyl makes: Mutant IDH, (R)-2-
hydroxyglutarate, and cancer. Genes Dev 27: 836–852.
Lowery DM, Lim D, Yaffe MB. 2005. Structure and function of Polo-like kinases. Oncogene 24: 248–
259.
Luo W, Hu H, Chang R, Zhong J, Knabel M, O’Meally R, Cole RN, Pandey A, Semenza GL. 2011.
Pyruvate kinase M2 is a PHD3-stimulated coactivator for hypoxia-inducible factor 1. Cell 145:
732–744.
MacKinnon R. 2003. Potassium channels. FEBS Lett 555: 62–65.
Malone JH, Oliver B. 2011. Microarrays, deep sequencing and the true measure of the transcriptome.
BMC Biol 9: 34.31–34.39.
Marks LR, Davidson BA, Knight PR, Hakansson AP. 2013. Interkingdom signaling induces
Streptococcus pneumoniae biofilm dispersion and transition from asymptomatic colonization to
disease. mBio 4: e00438-13.
Martell JD, Deerinck TJ, Sancak Y, Poulos TL, Mootha VK, Sosinsky GE, Ellisman MH, Ting AY.
2012. Engineered ascorbate peroxidase as a genetically encoded reporter for electron microscopy.
Nat Biotechnol 30: 1143–1148.
Martinez NJ, Gregory RI. 2010. MicroRNA gene regulatory pathways in the establishment and
maintenance of ESC identity. Cell Stem Cell 7: 31–35.
McGettigan PA. 2013. Transcriptomics in the RNA-seq era. Curr Opin Chem Biol 17: 4–11.
Mendell JT, Olson EN. 2012. MicroRNAs in stress signaling and human disease. Cell 148: 1172–1187.
Miller G. 2006. Optogenetics. Shining new light on neural circuits. Science 314: 1674–1676.
Miller EW, Lin JY, Frady EP, Steinbach PA, Kristan WB Jr, Tsien RY. 2012. Optically monitoring
voltage in neurons by photo-induced electron transfer through molecular wires. Proc Natl Acad Sci
109: 2114–2119.
Miller RM, Paavilainen VO, Krishnan S, Serafimova IM, Taunton J. 2013. Electrophilic fragment-
based design of reversible covalent kinase inhibitors. J Am Chem Soc 135: 5298–5301.
Miyawaki A. 2002. Green fluorescent protein-like proteins in reef Anthozoa animals. Cell Struct Funct
27: 343–347.
Miyawaki A. 2011. Development of probes for cellular functions using fluorescent proteins and
fluorescence resonance energy transfer. Annu Rev Biochem 80: 357–373.
Moffitt JR, Chemla YR, Smith SB, Bustamante C. 2008. Recent advances in optical tweezers. Annu
Rev Biochem 77: 205–228.
Monda JK, Scott DC, Miller DJ, Lydeard J, King D, Harper JW, Bennett EJ, Schulman BA. 2013.
Structural conservation of distinctive N-terminal acetylation-dependent interactions across a family
of mammalian NEDD8 ligation enzymes. Structure 21: 42–53.
Morgan DO. 2007. The cell cycle: Principles of control. New Science, London.
Moro E, Vettori A, Porazzi P, Schiavone M, Rampazzo E, Casari A, Ek O, Facchinello N, Astone M,
Zancan I, et al. 2013. Generation and application of signaling pathway reporter lines in zebrafish.
Mol Genet Genomics 288: 231–242.
* Morrison DK. 2012. MAP kinase pathways. Cold Spring Harb Perspect Biol 4: a011254.
Mulvihill MM, Nomura DK. 2013. Therapeutic potential of monoacylglycerol lipase inhibitors. Life Sci
92: 492–497.
Musacchio A, Noble M, Pauptit R, Wierenga R, Saraste M. 1992. Crystal structure of a Src-homology
3 (SH3) domain. Nature 359: 851–855.
Mutoh H, Perron A, Akemann W, Iwamoto Y, Knöpfel T. 2011. Optogenetic monitoring of membrane
potentials. Exp Physiol 96: 13–18.
Nardella C, Carracedo A, Salmena L, Pandolfi PP. 2010. Faithfull modeling of PTEN loss-driven
diseases in the mouse. Curr Top Microbiol Immunol 347: 135–168.
Nash PD. 2012. Why modules matter. FEBS Lett 586: 2572–2574.
Nash P, Tang X, Orlicky S, Chen Q, Gertler FB, Mendenhall MD, Sicheri F, Pawson T, Tyers M. 2001.
Multisite phosphorylation of a CDK inhibitor sets a threshold for the onset of DNA replication.
Nature 414: 514–521.
Newell EW, Sigal N, Nair N, Kidd BA, Greenberg HB, Davis MM. 2013. Combinatorial tetramer
staining and mass cytometry analysis facilitate T-cell epitope mapping and characterization. Nat
Biotechnol 31: 623–629.
Niu W, Guo J. 2013. Expanding the chemistry of fluorescent protein biosensors through genetic
incorporation of unnatural amino acids. Mol Biosyst 9: 2961–2970.
Nozawa YI, Lin C, Chuang PT. 2013. Hedgehog signaling from the primary cilium to the nucleus: An
emerging picture of ciliary localization, trafficking and transduction. Curr Opin Genet Dev 23:
429–437.
Otsubo Y, Yamamato M. 2008. TOR signaling in fission yeast. Crit Rev Biochem Mol Biol 43: 277–
283.
Overholtzer M, Mailleux AA, Mouneimne G, Normand G, Schnitt SJ, King RW, Cibas ES, Brugge JS.
2007. A nonapoptotic cell death process, entosis, that occurs by cell-in-cell invasion. Cell 131:
966–979.
Palumbo AM, Smith SA, Kalcic CL, Dantus M, Stemmer PM, Reid GE. 2011. Tandem mass
spectrometry strategies for phosphoproteome analysis. Mass Spectrom Rev 30: 600–625.
Parzych KR, Klionsky D. 2013. An overview of autophagy: Morphology, mechanism and regulation.
Antioxid Redox Signal 20: 460–473.
Patricelli MP, Szardenings AK, Liyanage M, Nomanbhoy TK, Wu M, Weissig H, Aban A, Chun D,
Tanner S, Kozarich JW. 2007. Functional interrogation of the kinome using nucleotide acyl
phosphates. Biochemistry 46: 350–358.
Pawson T, Nash P. 2003. Assembly of cell regulatory systems through protein interaction domains.
Science 300: 445–452.
Pearen MA, Muscat GE. 2012. Orphan nuclear receptors and the regulation of nutrient metabolism:
Understanding obesity. Physiology 27: 156–166.
* Perrimon N, Pitsouli C, Shilo B-Z. 2012. Signaling mechanisms controlling cell fate and embryonic
patterning. Cold Spring Harb Perspect Biol 4: a005975.
Piatkevich KD, Verkhusha VV. 2010. Advances in engineering of fluorescent proteins and
photoactivatable proteins with red emission. Curr Opin Chem Biol 14: 23–29.
Piatkevich KD, Subach FV, Verkhusha VV. 2013. Engineering of bacterial phytochromes for near-
infrared imaging, sensing, and light-control in mammals. Chem Soc Rev 42: 3441–3452.
Pluznick JL, Protzko RJ, Gevorgyan H, Peterlin Z, Sipos A, Han J, Brunet I, Wan LX, Rey F, Wang T,
et al. 2013. Olfactory receptor responding to gut microbiota-derived signals plays a role in renin
secretion and blood pressure regulation. Proc Natl Acad Sci 110: 4410–4415.
Portnoy DA. 2012. Yogi Berra, Forrest Gump, and the discovery of Listeria actin comet tails. Mol Biol
Cell 23: 1141–1145.
Pryciak PM. 2009. Designing new cellular signaling pathways. Chem Biol 16: 249–254.
Ptacek J, Snyder M. 2006. Charging it up: Global analysis of protein phosphorylation. Trends Genet
22: 545–554.
Racey LA, Byvoet P. 1971. Histone acetyltransferase in chromatin. Evidence for in vitro enzymatic
transfer of acetate from acetyl-coenzyme A to histones. Exp Cell Res 64: 366–370.
Rasmussen SG, DeVree BT, Zou Y, Kruse AC, Chung KY, Kobilka TS, Thian FS, Chae PS, Pardon E,
Calinski D, et al. 2011. Crystal structure of the β2 adrenergic receptor-Gs protein complex. Nature
477: 549–555.
Ravnskjaer K, Hogan MF, Lackey D, Tora L, Dent SY, Olefsky J, Montminy M. 2013. Glucagon
regulates gluconeogenesis through KAT2B- and WDR5-mediated epigenetic effects. J Clin Invest
123: 4318–4328.
Reményi A, Good MC, Lim WA. 2006. Docking interactions in protein kinase and phosphatase
networks. Curr Opin Struct Biol 16: 676–685.
Renicke C, Schuster D, Usherenko S, Essen LO, Taxis C. 2013. A LOV2 domain-based optogenetic
tool to control protein degradation and cellular function. Chem Biol 20: 619–626.
Resh MD. 2012. Targeting protein lipidation in disease. Trends Mol Med 18: 206–214.
* Rhind N, Russell P. 2012. Signaling pathways that regulate cell division. Cold Spring Harb Perspect
Biol 4: a005942.
Rockwell NC, Su YS, Lagarias JC. 2006. Phytochrome structure and signaling mechanisms. Annu Rev
Plant Biol 57: 837–858.
Rogne M, Taskén K. 2013. Cell signalling analyses in the functional genomics era. New Biotechnol 30:
333–338.
Roskoski RJ. 2012. MEK1/2 dual-specificity protein kinases: Structure and regulation. Biochem
Biophys Res Commun 417: 5–10.
Roux PP, Thibault P. 2013. The coming of age of phosphoproteomics—From large data sets to
inference of protein functions. Mol Cell Proteomics 12: 3453–3464.
Roy S, Hsiung F, Kornberg TB. 2011. Specificity of Drosophila cytonemes for distinct signaling
pathways. Science 332: 354–358.
Rubakhin SS, Lanni EJ, Sweedler JV. 2013. Progress toward single cell metabolomics. Curr Opin
Biotechnol 24: 95–104.
Sadikot RT, Blackwell TS. 2008. Bioluminescence: Imaging modality for in vitro and in vivo gene
expression. Methods Mol Biol 477: 383–394.
Sadowski MI, Taylor WR. 2013. Prediction of protein contacts from correlated sequence substitutions.
Sci Prog 96: 33–42.
Sadowsky JD, Burlingame MA, Wolan DW, McClendon CL, Jacobson MP, Wells JA. 2011. Turning a
protein kinase on or off from a single allosteric site via disulfide trapping. Proc Nat Acad Sci 108:
6056–6061.
Sakurai T, Amemiya A, Ishii M, Matsuzaki I, Chemelli RM, Tanaka H, Williams SC, Richardson JA,
Kozlowski GP, Wilson S, et al. 1998. Orexins and orexin receptors: A family of hypothalamic
neuropeptides and G protein-coupled receptors that regulate feeding behavior. Cell 92: 573–585.
Sanders TA, Llagostera E, Barna M. 2013. Specialized filopodia direct long-range transport of SHH
during vertebrate tissue patterning. Nature 497: 628–632.
Savinainen JR, Saario SM, Laitinen JT. 2012. The serine hydrolases MAGL, ABHD6 and ABHD12 as
guardians of 2-arachidonoylglycerol signalling through cannabinoid receptors. Acta Physiol 204:
267–276.
Scarpa ES, Fabrizio G, Di Girolamo M. 2013. A role of intracellular mono-ADP-ribosylation in cancer
biology. FEBS J 280: 3551–3562.
Schmid JA, Birbach A. 2007. Fluorescent proteins and fluorescence resonance energy transfer (FRET)
as tools in signaling research. Thromb Haemost 97: 378–384.
Schumacher MA, Chinnam N, Ohashi T, Shah RS, Erickson HP. 2013. The structure of irisin reveals a
novel intersubunit β-sheet fibronectin type III (FNIII) dimer: Implications for receptor activation. J
Biol Chem 288: 33738–33744.
Sengupta P, Van Engelenburg S, Lippincott-Schwartz J. 2012. Visualizing cell structure and function
with point-localization superresolution imaging. Dev Cell 23: 1092–1102.
Shaner NC, Lin MZ, McKeown MR, Steinbach PA, Hazelwood KL, Davidson MW, Tsien RY. 2008.
Improving the photostability of bright monomeric orange and red fluorescent proteins. Nat Methods
5: 545–551.
Shaw N, Liu ZJ. 2014. Role of the HIN domain in regulation of innate immune responses. Mol Cell
Biol 34: 2–15.
Shendure J, Mitra RD, Varma C, Church GM. 2004. Advanced sequencing technologies: Methods and
goals. Nat Rev Genet 5: 335–344.
Shu X, Royant A, Lin MZ, Aguilera TA, Lev-Ram V, Steinbach PA, Tsien RY. 2009. Mammalian
expression of infrared fluorescent proteins engineered from a bacterial phytochrome. Science 324:
804–807.
Shu X, Lev-Ram V, Deerinck TJ, Qi Y, Ramko EB, Davidson MW, Jin Y, Ellisman MH, Tsien RY.
2011. A genetically encoded tag for correlated light and electron microscopy of intact cells, tissues,
and organisms. PLoS Biol 9: e1001041.
Silverman JS, Skaar JR, Pagano M. 2012. SCF ubiquitin ligases in the maintenance of genome stability.
Trends Biochem Sci 37: 66–73.
Sims RJIII, Reinberg D. 2008. Is there a code embedded in proteins that is based on post-translational
modifications? Nat Rev Mol Cell Biol 9: 815–820.
Sopko R, Andrews BJ. 2008. Linking the kinome and phosphorylome—A comprehensive review of
approaches to find kinase targets. Mol Biosyst 4: 920–933.
Ståhl PL, Lundeberg J. 2012. Toward the single-hour high-quality genome. Annu Rev Biochem 81:
359–378.
Starai VJ, Takahashi H, Boeke JD, Escalante-Semerena JC. 2004. A link between transcription and
intermediary metabolism: A role for Sir2 in the control of acetyl-coenzyme A synthetase. Curr
Opin Microbiol 7: 115–119.
Stenkamp RE, Teller DC, Palczewski K. 2005. Rhodopsin: A structural primer for G-protein coupled
receptors. Arch Pharm 338: 209–216.
Strebhardt K. 2010. Multifaceted polo-like kinases: Drug targets and antitargets for cancer therapy. Nat
Rev Drug Discov 9: 643–660.
Stynen B, Tournu H, Tavernier J, Van Dijck P. 2012. Diversity in genetic in vivo methods for protein-
protein interaction studies: From the yeast two-hybrid system to the mammalian split-luciferase
system. Microbiol Mol Biol Rev 76: 331–382.
Tagliabracci VS, Xiao J, Dixon JE. 2013. Phosphorylation of substrates destined for secretion by the
Fam20 kinases. Biochem Soc Trans 41: 1061–1065.
Tang X, Orlicky S, Mittag T, Csizmok V, Pawson T, Forman-Kay JD, Sicheri F, Tyers M. 2012.
Composite low affinity interactions dictate recognition of the cyclin-dependent kinase inhibitor Sic1
by the SCFCdc4 ubiquitin ligase. Proc Natl Acad Sci 109: 3287–3292.
Tarrant MK, Cole PA. 2009. The chemical biology of protein phosphorylation. Annu Rev Biochem 78:
797–825.
Tartaglia LA, Dembski M, Weng A, Deng N, Culpepper J, Devos R, Richards GJ, Campfield LA,
Clark FT, Deeds J, et al. 1995. Identification and expression cloning of a leptin receptor, OB-R.
Cell 83: 1263–1271.
Thompson BJ. 2010. Developmental control of cell growth and division in Drosophila. Curr Opin Cell
Biol 22: 788–794.
Thorner J. 2006. Signal transduction. In Landmark papers in yeast biology (ed. Linder P, et al.), pp.
193–210. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY.
Tian X, Azpurua J, Hine C, Vaidya A, Myakishev-Rempel M, Ablaeva J, Mao Z, Nevo E, Gorbunova
V, Seluanov A. 2013. High-molecular-mass hyaluronan mediates the cancer resistance of the naked
mole rat. Nature 499: 346–349.
Toettcher JE, Weiner OD, Lim WA. 2013. Using optogenetics to interrogate the dynamic control of
signal transmission by the ras/erk module. Cell 155: 1422–1434.
Tong X, Shigetomi E, Looger LL, Khakh BS. 2013. Genetically encoded calcium indicators and
astrocyte calcium microdomains. Neuroscientist 19: 274–291.
Tsien RY. 2009. Constructing and exploiting the fluorescent protein paintbox. Angew Chem Int Ed
Engl 48: 5612–5626.
Turk BE, Cantley LC. 2003. Peptide libraries: At the crossroads of proteomics and bioinformatics. Curr
Opin Chem Biol 7: 84–90.
Turk BE, Hutti JE, Cantley LC. 2006. Determining protein kinase substrate specificity by parallel
solution-phase assay of large numbers of peptide substrates. Nat Protoc 1: 375–379.
Turner BM. 1991. Histone acetylation and control of gene expression. J Cell Sci 99: 13–20.
Urnov FD, Rebar EJ, Holmes MC, Zhang HS, Gregory PD. 2010. Genome editing with engineered zinc
finger nucleases. Nat Rev Genet 11: 636–646.
Vadas O, Burke JE, Zhang X, Berndt A, Williams RL. 2011. Structural basis for activation and
inhibition of class I phosphoinositide 3-kinases. Sci Signal 4: re.2.1–re.2.12.
Venta R, Valk E, Kõivomägi M, Loog M. 2012. Double-negative feedback between S-phase cyclin-
CDK and CKI generates abruptness in the G1/S switch. Front Physiol 3: 459.
Venter JC, Levy S, Stockwell T, Remington K, Halpern A. 2003. Massive parallelism, randomness and
genomic advances. Nat Genet 33: 219–227.
Vézina C, Kudelski A, Sehgal SN. 1975. Rapamycin (AY-22,989), a new antifungal antibiotic. I.
Taxonomy of the producing streptomycete and isolation of the active principle. J Antibiot (Tokyo)
28: 721–726.
Waksman G, Kominos D, Robertson SC, Pant N, Baltimore D, Birge RB, Cowburn D, Hanafusa H,
Mayer BJ, Overduin M, et al. 1992. Crystal structure of the phosphotyrosine recognition domain
SH2 of v-src complexed with tyrosine-phosphorylated peptides. Nature 358: 646–653.
Washietl S, Will S, Hendrix DA, Goff LA, Rinn JL, Berger B, Kellis M. 2012. Computational analysis
of noncoding RNAs. Wiley Interdiscip Rev RNA 3: 759–778.
Wei C, Liu J, Yu Z, Zhang B, Gao G, Jiao R. 2013. TALEN or Cas9—Rapid, efficient and specific
choices for genome modifications. J Genet Genomics 40: 281–289.
Weston CR, Davis RJ. 2001. Signaling specificity—A complex affair. Science 292: 2439–2440.
Wong SY, Reiter JF. 2008. The primary cilium at the crossroads of mammalian hedgehog signaling.
Curr Top Dev Biol 85: 225–260.
Woodward JJ, Iavarone AT, Portnoy DA. 2010. c-di-AMP secreted by intracellular Listeria
monocytogenes activates a host type I interferon response. Science 328: 1703–1705.
Wu H. 2013. Higher-order assemblies in a new paradigm of signal transduction. Cell 153: 287–292.
Wu H, Zhang Y. 2011. Mechanisms and functions of Tet protein-mediated 5-methylcytosine oxidation.
Genes Dev 25: 2436–2452.
Xiao H, Chatterjee A, Choi SH, Bajjuri KM, Sinha SC, Schultz PG. 2013a. Genetic incorporation of
multiple unnatural amino acids into proteins in mammalian cells. Angew Chem Int Ed Engl 52:
14080–14083.
Xiao J, Tagliabracci VS, Wen J, Kim SA, Dixon JE. 2013b. Crystal structure of the Golgi casein
kinase. Proc Natl Acad Sci 110: 10574–10579.
Xu J, Wang AH, Oses-Prieto J, Makhijani K, Katsuno Y, Pei M, Yan L, Zheng YG, Burlingame A,
Brückner K, et al. 2013. Arginine methylation initiates BMP-induced Smad signaling. Mol Cell 51:
5–19.
Yang W, Xia Y, Hawke D, Li X, Liang J, Xing D, Aldape K, Hunter T, Alfred Yung WK, Lu Z. 2012.
PKM2 phosphorylates histone H3 and promotes gene transcription and tumorigenesis. Cell 150:
685–696.
Zacharias DA, Tsien RY. 2006. Molecular biology and mutation of green fluorescent protein. Methods
Biochem Anal 47: 83–120.
Zhang Z, Tan M, Xie Z, Dai L, Chen Y, Zhao T. 2011. Identification of lysine succinylation as a new
post-translational modification. Nat Chem Biol 7: 58–63.
Zhang M, Galdieri L, Vancura A. 2013. The yeast AMPK homolog SNF1 regulates acetyl coenzyme A
homeostasis and histone acetylation. Mol Cell Biol 33: 4701–4717.
Zhou X, Herbst-Robinson KJ, Zhang J. 2012. Visualizing dynamic activities of signaling enzymes
using genetically encodable FRET-based biosensors from designs to applications. Methods Enzymol
504: 317–340.
Zimmermann H. 2008. ATP and acetylcholine, equal brethren. Neurochem Int 52: 634–648.
Zivanovic N, Jacobs A, Bodenmiller B. 2013. A practical guide to multiplexed mass cytometry. Curr
Top Microbiol Immunol doi: 10.1007/82_2013_335.
Zmajkovicova K, Jesenberger V, Catalanotti F, Baumgartner C, Reyes G, Baccarini M. 2013. MEK1 is
required for PTEN membrane recruitment, AKT regulation, and the maintenance of peripheral
tolerance. Mol Cell 50: 43–55.
Cite this chapter as Cold Spring Harb Perspect Biol doi: 10.1101/cshperspect.a022913
Index
A
ABI1, 186
AC. See Adenylyl cyclase
ACC. See Acetyl-CoA carboxylase
ACE. See Angiotensin-converting enzyme
Acetylation, protein regulation mechanisms, 34–35
Acetylcholine receptor, nicotinic, 267
Acetyl-CoA, 170–171, 176–177, 393–394, 434–435
Acetyl-CoA carboxylase (ACC), 287
ACL. See ATP-citrate lyase
Acrosome reaction, 336
ActA, 399
ACTH. See Adrenocorticotrophic hormone
Actin
cell migration and polymerization, 184–186
pathogen modifiers
elongation factors, 400
nucleation factors, 399–400
Activators of G-protein-mediated signaling (AGSs), 14
ADAM1b, 336
ADAM2, 336
ADAM3, 336
ADAM10, 111
ADAM17 (TACE), 111
Adenylyl cyclase (AC), 99–101, 335
Adiponectin, energy homeostasis role, 281
Adipose tissue triglyceride lipase (ATGL), 289
Adrenocorticotrophic hormone (ACTH), energy homeostasis role, 278
AF2, 130–131
AGS proteins, 14
AGSs. See Activators of G-protein-mediated signaling
AKAPs. See A-kinase anchoring proteins
A-kinase anchoring proteins (AKAPs), 43, 53–54, 101, 254, 335
Akt
cancer signaling, 408–416, 418–421
glucose metabolism signaling
G-protein-coupled receptor signaling, 13
lymphocyte signaling, 321–323
lymphocyte signaling, 322
mTORC1 target, 175
phosphoinositide, 3-kinase pathway overview, 87–89
PI3K/Akt, 168, 170–171
PIP3 signaling, 57–59
subcellular localization, 42
ALDH1A1, 335
α-Catenin, 20, 136
AMBRA1, 378
Amino acids, metabolism signaling cascades, 173, 175
2-Amino-3-(3-hydroxy-5-methyl-isoxazol-4-yl) propanoic acid receptor
(AMPAR), learning and memory role, 249–251, 253, 256
AMPAR. See, 2-Amino-3-(3-hydroxy-5-methyl-isoxazol-4-yl) propanoic
acid receptor
AMPK
energy homeostasis role, 278–279, 283–284, 287, 289
mTORC1 regulation, 176
Amyloid-β, 295
Anaphase. See Mitosis
Angiogenesis, cancer, 420
Angiomotin, 135
Angiotensin receptor, 273
Angiotensin-converting enzyme (ACE), 273
AP1, 318
APAF1, 368, 376–377
APC, 411
Apoptosis
cancer, 412
caspases
activation, function, and regulation, 367–368, 376
caspase-1/-5/-11 activation in inflammasome pathway, 374
caspase-2 activation in PIDDosome pathway, 374–375
caspase-8 activation in death receptor pathway, 371, 373–374
caspase-9 activation in mitochondrial pathway, 368–371
kinases, 375
overview, 366–367
TNFR1 induction, 303
Approximated, 134
Arp2/3, 186, 255, 305, 399–400
Arpp19, 155
ASC, 300, 374
ASK1, 351
ATF1, 101
ATF2, 349
ATF4, 349
ATF6, unfolded protein response, 347–349
ATG proteins. See Autophagy
ATM, 417
DNA damage checkpoint, 158–159
recruitment, 35
ATP-citrate lyase (ACL), 414
ATR, DNA damage checkpoint, 158–159
Aurora B, 41
Autophagy
cell death role, 379
signaling overview, 366, 377–379
B
Bacteria. See Infection
Bad, 372, 378
BAFF, 319–320, 322
Bak, 369
Bax, 369
B-cell receptor (BCR)
adaptor molecules, 316
ITAM, 315–316
signaling
calcium, 316–317
diacylglycerol, 316–318
ERK1/2, 319
inhibitory signals, 323
nuclear factor of activated T cells, 317–318
overview, 125–127, 317
protein kinase C, 318–319
Ras, 319
structure and function, 314–315
Bcl2, 356, 367, 369–371, 378–379, 412
BCL6, 334
Bcl10, 307, 319
BCR. See B-cell receptor
BCR-ABL, 407
Beclin, 1, 378–379
BEN. See Biased excitable network
β-Arrestin, G-protein-coupled receptor complex, 14, 107
β-Catenin, 20, 104, 417
β-TrCP, 370
Biased excitable network (BEN), 193
Bid, 370, 372
BIK, 372
Bim, 356, 370, 372, 412
BK channel, 62
BLNK, 316
Bmf, 356, 373
BMP. See Bone morphogenetic protein
Bnip3, 373, 378
BOC, 107
Bone morphogenetic protein (BMP)
BMPRII, 114
embryonic patterning, 223, 225
signaling overview, 113–114
BRCA, 418
Brinker, 226
C
Cadherins, signaling, 20
Calcineurin, learning and memory role, 253–254
Calcium
binding motifs, 61
buffering, 61–62
channels and regulation of levels, 61
history of study, 95–96
lymphocyte signaling, 316–317
signaling overview, 59–61, 95–97
smooth muscle sensitization, 273
spatiotemporal organization of signaling, 62
termination of signal, 62
Calcium/calmodulin-dependent protein kinase II (CaMKII), learning and
memory role, 251–253, 258
Calcium-induced calcium release (CICR), 62, 269
Calcium release-activated channel (CRAC), 305, 316–317
Calmodulin (CaM), calcium signaling, 97
Calsequestrin (CSQ), 268–269, 271
CaM. See Calmodulin
CAMKII, caspase activation role, 375
CaMKII. See Calcium/calmodulin-dependent protein kinase II
cAMP. See Cyclic AMP
CAMs. See Cell adhesion molecules
Cancer
cell polarity signaling, 210
dysregulation
angiogenesis, 420
apoptosis, 412
cell fate and differentiation, 417
cell migration, 415–417
cell polarity, 415–417
cell proliferation, 409–412
extracellular matrix, 418
genomic instability, 417–418
inflammation, 420–421
metabolism, 412–415
microenvironment, 418
fibroblasts, 421
gene mutations
overview, 407
signaling pathways, 407–408, 419
inflammation, 307
progression, 406–407
prospects for study, 421–422
CAR. See Constitutive active/androstane receptor
Carbohydrate. See Glucose, metabolism signaling
Carbon monoxide (CO), signal transduction, 25
Carboxy-terminal Src kinase (CSK), 315
Cardiac muscle. See Muscle contraction
Carma1, 319
Carnitine:palmitoyl transferase (CPT), 284
Casein kinase II (CK2), 370
Caspases. See Apoptosis; Necrosis
Catecholaminergic polymorphic ventricular tachycardia (CPVT), 271
CATSPER, 336
Caveolae, signaling, 43
Cbl, 8
Cdc2, 146, 153
Cdc4, 145
Cdc13, 146
Cdc20, 148, 160–161
Cdc25, 155, 157–159, 330
Cdc34, 146
Cdc42, 82, 354, 368, 98
cell migration role, 186, 188
cell polarity role, 200–201, 207
CDKs. See Cyclin-dependent kinases
CDO, 107
Cdr2, 157
C/EBPα, 131
C/EBPβ, 297
CED3, 376
CED4, 376
CED9, 376
Cell adhesion molecules (CAMs). See also specific molecules
cadherin-dependent adhesions, 20
signaling, 19–21
Cell cycle
cancer, 409–412
cyclin-dependent kinase inhibitors
overview, 143–144
transcriptional regulation by inhibitors, 144–145
G1 regulation
cyclin D, 141–142
cyclin E, 142–143
cyclin-dependent kinases, 141, 143
retinoblastoma protein, 140–141
ubiquitinylation, 145–147, 147–148
G2/M transition. See Mitosis
meiosis. See Meiosis
overview, 140–141
Cell fate. See Cancer; Embryonic patterning, Drosophila
Cell migration
actin polymerization, 184–186
adhesions, 186–188
cancer, 415–417
chemotaxis signaling
adaptation and excitation–global inhibition models, 193
excitability of networks, 192–193
myosin contraction, 186
overview, 184
polarization, 188
prospects for study, 193–195
signaling
focal adhesion kinase, 188–189
genetic analysis, 191–192
paxillin, 188–189
phosphoinositide, 3-kinase, 189–191
Rho GTPases, 188
Cell polarity. See also Par proteins
cancer, 415–417
cell migration, 188
machinery
intercellular junctions, 203
Par proteins, 202–203
symmetry breaking and positive-feedback loops, 200–202
Par protein localization
active exclusion, 205–207
membrane phospholipid attachment, 204
membrane protein anchoring, 204–205
messenger RNA localization, 205
oligomerization, 204
signaling
cancer, 210
Hippo pathway, 209–210
overview, 200
Par3–Par6–protein kinase C signaling, 207–209
Wnt signaling cross talk, 209
Ceramide
hydrolysis, 59
signaling overview, 56
stress response, 59
cGMP. See Cyclic GMP
CHBP, 400
Chemotaxis. See Cell migration
Chk1, 158–159, 417
Chk2, 158–159, 417
Cholecystokinin, receptor, 273
Cholera toxin, 390
Cholesterol
Hedgehog coupling, 107
liver metabolism, 287–289
CHOP, 350, 352
CICR. See Calcium-induced calcium release
Cif, 400
CIN85, 316
CK1, 317
CKII. See Casein kinase II
CKS1, 147
Cks1, 434
Clb5, 146, 434
Cln2, 434
CO. See Carbon monoxide
Colony-stimulating factor, 1 (CSF1), 420
Complement, C5A receptor in innate immunity, 304–305
Computational models, signaling networks
dynamical models, 71–72
graph theory for signaling network models, 69, 72–74
network models, 70–71
Constitutive active/androstane receptor (CAR), 23–24
Cortisol, energy homeostasis role, 278–279
Cos, 107
COX2, 297
CPI-17, 273
CPT. See Carnitine:palmitoyl transferase
CPVT. See Catecholaminergic polymorphic ventricular tachycardia
CRAC. See Calcium release-activated channel
Crb2, 159
CRE. See Cyclic AMP response element
CREB, 274, 297, 334
CRIB domain, 398
Crk, 8
CRL4, 148
Crumbs, 135, 204, 416
CSF. See Cytostatic factor
CSF1. See Colony-stimulating factor, 1
CSK. See Carboxy-terminal Src kinase
CSL complex, 110–111
CSQ. See Calsequestrin
CTLA4, 319, 323
Ctp1, 158
Cyclic AMP (cAMP)
G-protein-coupled receptor signaling, 13, 53
muscle relaxation, 265
phosphodiesterases, 55
protein kinase A as target, 53–54
signaling overview, 99–101
targets, 55
Cyclic AMP response element (CRE), 287
Cyclic GMP (cGMP)
muscle relaxation, 265
phosphodiesterases, 55
protein kinase G as target, 55
targets, 55
Cyclic GMP-dependent protein kinase (PKG), 55
Cyclin B, mitosis entry role, 154–156
Cyclin D
G1 regulation, 141–142
ubiquitinylation, 145–147
Cyclin-dependent kinases (CDKs)
activating kinase, 154
CDK1
activation in mitosis entry, 152–157
caspase activation, 375–377
oocyte maturation role, 331
G1 entry regulation, 141
inhibitors
overview, 143–144
transcriptional regulation
INK4, 144–145
p21, 144
posttranscriptional regulation, 143
Cyclin E, G1 regulation, 142–143
Cyclophilin D (CypD), 380
Cyclosporin A, 318
CYLD, 380
CypD. See Cyclophilin D
Cytochrome c, 376
Cytokine receptor family, overview, 4, 6
Cytokinesis, mitosis coordination, 161–162
Cytostatic factor (CSF), 328
D
Dachsous, 134
DAI, 380
Dcn1, 434
DCP1, 376
Death receptors. See specific receptors
Deltex, 110
Development. See Embryonic patterning, Drosophila
Diabetes type, 2, overnutrition, 289–290
Diacylglycerol (DAG)
lymphocyte signaling, 316–318
muscle calcium sensitization, 273
protein kinase C as target, 55, 57
signaling overview, 57
DIAP1, 376
DISC, 374
Discs large, 135
DKK, 104
Dlg, 204
DLG1, 57
DNA damage checkpoint, 157–159
DNA-PK, DNA damage checkpoint, 158
DNA replication checkpoint, 159
DNMT1, 407
DOCK180, 19
Dock2, 305
Dorsal, 217
Double-strand break (DSB), 157–158
Double-stranded RNA-dependent kinase (PKR), 349, 351
DrrA, 398
DSB. See Double-strand break
DUSP, 355
E
E2F, 141, 143
E4orf6, 401
E-cadherin, 20
Ect1, 207
EGF. See Epidermal growth factor
eIF2, 349
Electron microscopy, signaling studies, 430
Elk1, 40
Embryonic patterning, Drosophila
induction of cell fate, 216
signaling
bone morphogenetic protein, 223, 225
epidermal growth factor, 222–224, 229
fibroblast growth factor, 222–223, 229
Hedgehog, 222, 225
integration of pathways, 227–230
linear signaling, 226
long-range ligand distribution, 225–226
negative-feedback switches, 226
Notch, 220–222
overview, 216–218
threshold generation, 226–227
Wnt, 225
transcriptional cascade interactions with signaling, 217, 219
Emi2, 331, 338
EMT. See Epithelial-to-mesenchymal transition
Endoplasmic reticulum. See Unfolded protein response
Endoplasmic reticulum stress element (ERSE), 347
Endosomal sorting complex required for transport (ESCRT), 9, 111
Energy homeostasis. See also Fatty acid metabolism; Glucose, metabolism
signaling
AMPK role, 278–279, 283–284, 287, 289
brown adipose tissue fatty acid oxidation, 289
diabetes type, 2 and insulin resistance, 289–290
hormonal control
adipose tissue, 281
adrenal gland, 278
hypothalamic-pituitary axis, 278
pancreas, 280
thyroid gland, 279–280
liver
carbohydrate metabolism acute regulation, 284–286
gluconeogenesis long-term regulation, 286–287
lipid metabolism, 287–289
muscle
exercise adaptation, 284
fatty acid oxidation, 284
glucose uptake and glycogen synthesis, 282–284
glycogen breakdown, 281–282
prospects for study, 290–291
ENTPD5, 168
EphB6, 8
Epidermal growth factor (EGF)
embryonic patterning, 222–224, 229
receptor
cancer, 9
dimerization, 7
phosphorylation, 433
Epithelial-to-mesenchymal transition (EMT), cancer, 416–417
EPSP. See Excitatory postsynaptic potential
ERK. See Extracellular signal-regulated kinase
ERO1, 349
ERR. See Estrogen-related receptor
ERSE. See Endoplasmic reticulum stress element
ESCRT. See Endosomal sorting complex required for transport
E-selectin, innate immunity role, 303–304
EspG, 398
Estrogen-related receptor (ERR), 21, 23
Ets, 334
Excitatory postsynaptic potential (EPSP), 248
ExoS, 396
Extracellular signal-regulated kinase (ERK)
cancer signaling, 408–410, 412, 415–416, 418, 421
development role, 223, 226–227, 230
learning and memory role, 258
lymphocyte signaling, 319
lymphocyte signaling by ERK1/2, 319
overview, 81–82
stress signaling
activation cascade, 354–355
inactivation, 355
overview, 353–354
physiological roles
cell death, 356
inflammation, 356–357
metabolism, 357
scaffold protein function, 355–356
substrates, 41
F
FADD, 16, 300, 302–303, 373–374
FAK. See Focal adhesion kinase
Far1, 146
Farnesoid X receptor (FXR), 23
Fas, 15
Fat, 134
Fatty acid metabolism
brown fat oxidation, 289
muscle oxidation, 284
FBW7, 146
Fbxo proteins, 146
Fc receptors
FcεR1 and immunoglobulin E binding, 316
innate immunity
FcεRI, 305–307
FcγR, 307
Fertilization. See Reproduction
FGF. See Fibroblast growth factor
Fibroblast growth factor (FGF)
cancer dysregulation, 417
embryonic patterning, 222–223, 229
receptor
dimerization, 7
stabilization, 40
FIDOP domain, 397
FIP200, 379
FK506, 318
FKBP12, 430
FLIP, 303, 373–375, 380, 412
Fluorescence resonance energy transfer (FRET), signaling studies, 429
FNDC5, 437
Focal adhesion kinase (FAK), 18, 20, 188–189, 418
Follicle-stimulating hormone (FSH), 334–335
Formyl peptide receptor (FPR), innate immunity, 304–305
4E-BP1, 44
Fos, 41, 274
Foxo
cancer signaling, 414–415
FOXO1, 87
FOXO3A, 371
lymphocyte signaling, 322
FPR. See Formyl peptide receptor
FRET. See Fluorescence resonance energy transfer
Frizzled, 10, 103, 209
FSH. See Follicle-stimulating hormone
FUNDC1, 379
Fus3, 146
FXR. See Farnesoid X receptor
Fyb, 392
Fyn, 20
G
G protein
GTPase cycle, 37
heterotrimeric G proteins, 38–39
receptor specificity, 11–12
small G proteins, 37–38
G-protein-coupled receptors (GPCRs)
activation, 11
classes and structure, 10–11
G-protein specificity, 11–12
guanine nucleotide exchange factors, 37, 39
innate immunity, 304–305
kinase networks, 13–14
ligand-induced conformational change, 12–13
lipid messengers, 57
overview, 10
pathogen signaling corruption in host
bacteria protein mimics
G proteins, 398
GTPase-activating proteins, 396
guanine-nucleotide exchange factors, 395–395
G-protein modifiers, 397–398
overview, 395
Yersinia, 396–397
protein–protein interactions, 14
second messengers, 13
sensory receptors. See Sensory receptors
signal integration, 15
signal termination, 14–15
G1. See Cell cycle
G2/M transition. See Mitosis
GADD34, 350
GADD45, 355
GATA4, 269
Gcn5, 176, 284
GDNF. See Glial-derived neurotrophic factor
Genomic instability, cancer, 417–418
Germ-cell nuclear receptor, 24
Germinal vesicle breakdown (GVBD), 330
GFP. See Green fluorescent protein
GFRA1, 334
GH. See Growth hormone
Gli1, 107–108
Gli2, 107
Gli3, 107
Glial-derived neurotrophic factor (GDNF), 8, 333–334
Glucocorticoid response element (GRE), 286
Glucose
liver
carbohydrate metabolism acute regulation, 284–286
gluconeogenesis long-term regulation, 286–287
metabolism signaling
cancer, 412–415
endoplasmic reticulum signals, 351–353
hypoxia-inducible factor, 1, 171–173
PI3K/Akt, 168, 170–171
pyruvate kinase metabolic switch, 173–174
muscle uptake and glycogen synthesis, 282–284
transporters, 168, 282, 284
Glycogen, muscle
breakdown, 281–282
glucose uptake and synthesis, 282–284
Glycosylation, protein regulation mechanisms, 36–37
GPCRs. See G-protein-coupled receptors
GPR3, 329
Graph theory, signaling network models, 69, 72–74
Grb2, 18, 320
GRE. See Glucocorticoid response element
Green fluorescent protein (GFP), signaling studies, 428
Growth hormone (GH), receptor dimerization, 7
GRP78, 347, 351
GRP94, 347
GSK3, 104, 145–146, 284, 317, 370, 411, 433
Guanylyl cyclase, 24–25
Gustation. See Sensory receptors
GVBD. See Germinal vesicle breakdown
H
H2AX, 158–159
HAT. See Histone acetyltransferase
HCN. See Hyperpolarization-activated cyclic nucleotide-gated channel
HDAC. See Histone deacetylase
Hearing. See Sensory receptors
Heart failure. See Muscle contraction
Hedgehog (Hh)
embryonic patterning, 222–223, 225
signaling overview, 107–108
Heme oxygenase (HO), 25
Hemese, embryonic patterning, 219–220
HES1, 227
Hexokinase, 168
Hh. See Hedgehog
HIF1. See Hypoxia-inducible factor, 1
Hippo
cell polarity signaling, 209–210
signaling overview, 131–136
Histone acetyltransferase (HAT), 35
Histone deacetylase (HDAC), 35, 130, 268
HMG-CoA reductase, 287–288
HMGB1, 294
HNF4, 131
Hormone-sensitive lipase (HSL), 289
HRK, 373
HSL. See Hormone-sensitive lipase
HuR, 356
2-Hydroxyglutarate aciduria, 176–177
Hyperpolarization-activated cyclic nucleotide-gated channel (HCN), 55
Hypoxia-inducible factor, 1 (HIF1)
glucose metabolism signaling, 171–171
oxygen signaling, 24
I
ICAM. See Intercellular cell adhesion molecule
ICP0, 400–401
IDH. See Isocitrate dehydrogenase
IKK, 45, 123, 297–299, 303, 307, 351, 393
induced genes, 298
Toll-like receptor activation, 295–297
IL-2 receptor, 320–321
IL-3
metabolism regulation, 176
receptor, 176
IL-4 receptor, 321
IL-7 receptor, 322
IL-12 receptor, 320
ILK. See Integrin-linked kinase
Immunoreceptor tyrosine activation motif (ITAM), 303, 315–317
Immunoreceptor tyrosine inhibition motif (ITIM), 323
Infection
actin modifiers
elongation factors, 400
nucleation factors, 399–400
bacteria virulence factors, 390–392
G-protein-coupled receptor signaling corruption
overview, 395
bacteria protein mimics
G proteins, 398
GTPase-activating proteins, 396
guanine-nucleotide exchange factors, 395–395
G-protein modifiers, 397–398
Yersinia, 396–397
lipid signaling highjacking, 398–399
mitogen-activated protein kinase signaling corruption
anthrax, 393
overview, 392
Shigella, 394
Yersinia, 393–394
oncoproteins of viruses, 390
ubiquitylation disruption, 400–401
Inflammation
cancer, 307, 420–421
caspase-1/-5/-11 activation in inflammasome pathway of apoptosis, 374
mitogen-activated protein kinase stress signaling, 356–357
unfolded protein response, 350–351
Information flow, signaling networks
computational models. See Computational models, signaling networks
contextual nature of information, 75
emergent properties of networks
bistability, 74
oscillations, 75
redundancy, 75
ultrasensitivity, 74
graph theory models, 69, 72–74
networks of pathways, 66, 68–69
noise filtering, 76
overview, 66–67
versatility of responses, 75–76
INK4, transcriptional regulation, 144–145
Innate immunity. See also specific receptors
E-selectin, 303–304
Fc receptors
FcεRI, 305–307
FcγR, 307
G-protein-coupled receptors, 304–305
inflammation and cancer, 307
integrin receptors, 303–304
Nod-like receptors, 300
pattern recognition receptor ligands, 294–295
RIG-I-like receptors, 298–300
TNFR1 signaling
cell death induction, 303
MAPK, 301–303
nuclear factor-κB, 301–303
Toll-like receptor
interferon induction, 297–298
ligands, 295
TAK1 and IKK activation, 295–296
TLR4 signaling, 296
TRIF in signaling, 298
Inositol hexaphosphate (IP6), 57
Inositol tetraphosphate (IP4), 57
Inositol-requiring enzyme, 1 (IRE1), unfolded protein response, 347–348,
350
Inositol trisphosphate (IP3)
G-protein-coupled receptor signaling, 13, 57, 96
receptors, 60, 273, 337
Insulin
energy homeostasis role, 280
resistance and overnutrition, 289–290
response unit, 287
Insulin receptor substrate (IRS), 87
Integrin receptors
innate immunity, 303–304
ligands, 18
mechanosensing signaling, 17–19
Integrin-linked kinase (ILK), 18–19, 189
Intercellular cell adhesion molecule (ICAM), 17, 303
IP3. See Inositol trisphosphate
IP4. See Inositol tetraphosphate
IP6. See Inositol hexaphosphate
IpgD, 398–399
IRAK1, 123, 295, 298
IRE1. See Inositol-requiring enzyme, 1
IRF1, 298
IRF7, 297–298, 401
IRS. See Insulin receptor substrate
IRS1, 351
Ishihara test, 243
Isocitrate dehydrogenase (IDH), 176, 178–179, 407, 415, 436
ITAM. See Immunoreceptor tyrosine activation motif
ITIM. See Immunoreceptor tyrosine inhibition motif
IZUMO, 336
J
JAK/STAT signaling
cytokine signaling in lymphocytes, 320–321
overview, 117–119
JAM. See Junctional adhesion molecule
Janus kinase. See JAK/STAT signaling
JIP1, 356
JIP2, 356
JIP3, 356
JIP4, 356
JNK. See Jun N-terminal kinase
Jun N-terminal kinase (JNK)
overview, 82–83
stress signaling
activation cascade, 354–355
inactivation, 355
overview, 353–354
physiological roles
cell death, 356
inflammation, 356–357
metabolism, 357
scaffold protein function, 355–356
Junctional adhesion molecule (JAM), 204–205
K
KIP1, 147
Kit, 9, 334
KitL. See Stem cell factor
KSR1, 356
Ku70, 158
Ku80, 158
L
Lactate dehydrogenase (LDH), 414
LAMP1, 378
LAMP2, 378
LAT, 126, 316
LATS kinases, 162
LC3, 379
Lck, 315
LDH. See Lactate dehydrogenase
Learning and memory
GTPases in synaptic plasticity, 257–259
Hebbian behavior of synapses, 248–249
postsynaptic density scaffold proteins, 254–256
prospects for study, 259
spine synapse signaling in brain
2-amino-3-(3-hydroxy-5-methyl-isoxazol-4-yl) propanoic acid
receptor, 249–251, 253, 256
calcium-regulated signaling in postsynaptic density
calcium/calmodulin-dependent protein kinase II, 251–253, 258
calcineurin, 253–254
N-methyl-D-aspartate receptor, 249–251
overview, 249–251
LEF/TCF, 114
LEGI model, 193
Leptin, 281
Lethal factor, anthrax, 393
Leukotrienes, signaling overview, 59
Leydig cell, 332
Lgl, 204–205, 210
LGP2, 298–299
LH. See Luteinizing hormone
LIM kinase, 186
Linear ubiquitin chain assembly complex (LUBAC), 295, 303
Lipid raft, signaling, 43
Lipids. See Fatty acid metabolism; specific lipids
Liver X receptor (LXR), 23, 130–131
LKB1, 415
LMP1, 407
Long-term depression (LTD), 249, 251–253
Long-term potentiation (LTP), 249
Lowfat, 134
LRP4, 8
LRP6, 228
LTD. See Long-term depression
LTP. See Long-term potentiation
LUBAC. See Linear ubiquitin chain assembly complex
Luteinizing hormone (LH), 334
LXR. See Liver X receptor
M
MAD, 134, 227
Magnesium
signaling overview, 59–60, 62–63
transport, 63
MAGUKS, 254
MALT1, 307, 319, 420
MAM. See Mitochondria-associated endoplasmic reticulum membrane
MAML, 111
Mammalian target of rapamycin (mTOR)
complexes. See mTORC1; mTORC2
G-protein-coupled receptor signaling, 13
history of study, 92–93
MAPK. See Mitogen-activated protein kinase
Maturation promoting factor (MPF), 328, 330
MCC. See Mitotic checkpoint complex
Mcl1, 356, 369–370, 379, 412
Mdc1, 158
mDia1, 186, 305
mDia2, 186
MDM2, 43–44, 411
Mechanosensation. See Sensory receptors
MEF2, 219, 268
Meiosis
oocyte
meiosis I, 328–331
meiosis II arrest, 331–332
spermatocyte meiosis and release, 335
Memory. See Learning and memory
MEN. See Mitotic exit networks
Messenger RNA (mRNA), stability, 44
Met1, 434
Methylation, protein regulation mechanisms, 35–36
MGA5, 298–300
MHCK. See Myosin heavy-chain kinase
MicroRNA, signaling, 44
Microtubule-organizing center (MTOC), 188, 318
Migration. See Cell migration
Mitochondria-associated endoplasmic reticulum membrane (MAM), 346–347
Mitochondrial outer membrane permeabilization (MOMP), 368–371, 376
Mitogen-activated protein kinase (MAPK). See also specific kinases
activation, 35
cascade, 81
classification, 81–83
computational models, 72–73
cyclin D transcription control, 141
development role, 217
embryonic patterning, 217
G-protein-coupled receptor signaling, 13
interacting kinases, 410–411
learning and memory role, 258
motifs, 40
pathogen signaling corruption in host
anthrax, 393
overview, 392
Shigella, 394
Yersinia, 393–394
scaffolds, 43, 85
stress signaling
activation cascade, 354–355
inactivation, 355
overview, 353–354
physiological roles
cell death, 356
inflammation, 356–357
metabolism, 357
scaffold protein function, 355–356
subcellular localization, 41
synaptic plasticity role, 258
TNFR1 signaling, 301–303
unfolded protein response, 346
Mitosis
anaphase entry, 160
CDK1 activation, 152–157, 160
cytokinesis coordination, 161–162
DNA damage checkpoint, 157–159
DNA replication checkpoint, 159
G2/M transition regulation, 157
network dissection, 160
spindle assembly checkpoint, 160–161
table of proteins in control, 154
transitions, 152–153
Mitotic checkpoint complex (MCC), 161
Mitotic exit network (MEN), 161–162
MLCK. See Myosin light-chain kinase
MLCP. See Myosin light-chain phosphatase
MLKL, 303, 380
MMSET, 36
Modularity, signaling studies, 431–432
MOMP. See Mitochondrial outer membrane permeabilization
Mos, 330–331
M-phase promoting factor (MPF), 153
MPF. See Maturation promoting factor; M-phase promoting factor
Mrc1, 150
MRN complex, 158
mRNA. See Messenger RNA
MscC, 238
MscL, 238–239, 245
MscS, 238–239
MSK1, 303, 394
MSK2, 394
MTOC. See Microtubule-organizing center
mTOR. See Mammalian target of rapamycin
mTORC1, 88, 174–175
activation and regulation, 88, 92–93, 176
metabolic signaling, 178
translational control, 44
unfolded protein response studies, 350
mTORC2, 92–93, 175
Muscle contraction
cardiac muscle
contraction, 269
heart failure, 271
hypertrophy
exercise-induced, 269–270
pathophysiological, 270–271
energy homeostasis
exercise adaptation, 284
fatty acid oxidation, 284
glucose uptake and glycogen synthesis, 282–284
glycogen breakdown, 281–282
signaling overview, 264–266
skeletal muscle
contraction, 266–268
fiber types and exercise response, 268–269
malignant hyperthermia, 269
smooth muscle
calcium sensitization, 273
contraction, 272–273
types, 271
vascular disease, 273–274
Myc, 41, 173, 408, 411, 415, 418
MyD88, 123, 295–298, 300
Myopic, 134
Myosin
cell migration and contraction, 186
myosin II, 186
Myosin heavy-chain kinase (MHCK), 191
Myosin light-chain kinase (MLCK), 186, 265, 268, 273, 304
Myosin light-chain phosphatase (MLCP), 265, 273
MYPT1, 273
Myristoylation, membrane proteins, 42
Myt1, 330
N
NBR1, 379
Nbs1, 158
NCAM. See Neural cell adhesion molecule
Nck, 8, 189
NCS1, 271
Necrosis
caspase control, 380
excitotoxicity, 381
Nox1 induction, 380–381
overview, 366
types, 379–380
Nedd4, 110, 113
Nedd8, 45
Nek2, 41
NEMO, 45, 123
Nerve growth factor (NGF), receptor dimerization, 7
Neural cell adhesion molecule (NCAM), signaling, 20–21
Neuromuscular junction (NMJ), 267
NF-κB. See Nuclear factor-κB
NF1, 411
NF2, 134
NFAT. See Nuclear factor of activated T cells
NGF. See Nerve growth factor
Nitric oxide (NO)
muscle relaxation, 265
signal transduction, 24–25, 37
Nitrosylation, protein regulation mechanisms, 37
NIX, 373, 379
NLR. See Nod-like receptor
NLRC4, 300, 303
NLRP1, 300, 303
NLRP3, 300, 303, 351
NMDAR. See N-Methyl-D-aspartate receptor
N-Methyl-D-aspartate receptor (NMDAR), learning and memory role, 249–
251
NMJ. See Neuromuscular junction
NO. See Nitric oxide
Nod-like receptor (NLR), signaling, 300
NOD1, 300
NOD2, 300
Noise filtering, signaling networks, 76
Notch
development role, 220–221
embryonic patterning, 220–222
intracellular domain, 111–112, 220
proteolysis and activation, 9–10
signaling overview, 109–111
Nox1, necrosis induction, 380–381
Noxa, 372
NPAS2, 25
NR4A receptors, 24
NR5A receptors, 24
NRF1, 284
NRF2, 284
NSK1, 303
Nuclear factor of activated T cells (NFAT), 127, 268, 274, 307, 317–318, 416
Nuclear factor-κB (NF-κB)
induced genes, 298
lymphocyte signaling, 127
TLR signaling, 297
TNFR1 signaling, 301–303
Nuclear receptors. See also specific receptors
activation, 21
classification, 21–22
orphan receptors, 24
overview, 21, 129–132
promoter binding
heterodimers, 22
homodimers, 22
monomers, 22–23
structure, 21–22, 130
types, 23–24, 130
Numb, cell polarity role, 207–207
O
Oct4, 173
Odd paired, embryonic patterning, 219
Olfaction. See Sensory receptors
Omi, 369
Oocyte. See also Reproduction
activation on fertilization, 336–338
maturation
meiosis I, 328–331
meiosis II arrest, 331–332
overview, 328
OSM, 356
OspF, 394
Oxidative stress, unfolded protein response, 349
Oxygen, signal transduction, 24
P
p16, 144–145, 411
p18, 144–145
p21, 144, 143–144, 146–148, 411, 417
p27, 144, 411
p38 stress-activated protein kinase (SAPK)
cell cycle checkpoint, 159
overview, 84–85
stress signaling
activation cascade, 354–355
inactivation, 355
overview, 353–354
physiological roles
cell death, 356
inflammation, 356–357
metabolism, 357
scaffold protein function, 355–356
p53, 43–44, 417, 420
p57, 144, 147
PAK, 113, 189, 258, 398
PAK2, 368
PAK3, 193
Palmitoyl acyltransferase (PAT), 42
Palmitoylation, membrane proteins, 42–43
Pals1, 204
PAR1. See Protease-activated receptor, 1
Par proteins
cell polarity role, 202–203
localization
active exclusion, 205–207
membrane phospholipid attachment, 204
membrane protein anchoring, 204–205
messenger RNA localization, 205
oligomerization, 204
Par1, 205
Par2, 202
Par3, 202–203, 205–208
Par3–Par6–protein kinase C signaling, 207–209
Par5, 205
Par6, 202–208
Parvin, 189
PAT. See Palmitoyl acyltransferase
Patched, 107–108
Pathogens. See Infection
Patj, 204, 210
PAX2, 229
Paxillin, cell migration role, 188–189
PCNA, 411
PD-1, 319
PDEs. See Phosphodiesterases
PDGF. See Platelet-derived growth factor
PDK. See Pyruvate dehydrogenase kinase
PEA15, 355
Peli1, 298
PEPCK. See Phosphoenolpyruvate carboxykinase
PERK, 347–348, 350, 352
Perlipin1, 289
Permeability transition pore (PTP), 63
Peroxisome proliferator-activated receptor (PPAR), 23, 132
PFK. See Phosphofructokinase
PGC1α, 268, 270, 274, 284
PHAPI, 371
PHD. See Prolyl hydroxylase
Phosphatidylinositol bisphosphate (PIP2)
bacteria hydrolysis, 398–399
signaling overview, 55–57, 63
Phosphatidylinositol trisphosphate (PIP3)
Akt signaling, 57–59, 87
lymphocyte signaling, 322–323
Phosphodiesterases (PDEs), 99
oocyte PDE3, 329
overview, 55
Phosphoenolpyruvate carboxykinase (PEPCK), 286–287
Phosphofructokinase (PFK), 168, 285–286
Phosphoinositide, 3-kinase (PI3K)
activation, 52, 58, 87
Akt pathway overview, 87–89
cancer signaling, 408–416, 418–421
cell migration role, 189–191
G-protein-coupled receptor signaling, 13–14
glucose metabolism signaling, 168, 170–171
lymphocyte signaling, 126–127
lymphocyte signaling, 321–323
mTORC1 target, 175
recruitment, 42
subcellular localization, 42
Phospholamban, 269
Phospholipase A2 (PLA2), isoforms, 432–433
Phospholipase C (PLC)
activation, 52
FcR signaling, 307
feedback control, 9
fertilization role, 337–338
LAT recruitment, 316
lymphocyte signaling, 126
messenger generation, 56–57
muscle calcium sensitization, 273
Phosphorylation, protein regulation mechanisms, 33–34
PI3K. See Phosphoinositide, 3-kinase
PIDD, 374–375
PIF, 429
Pins, cell polarity role, 205–206
PIP2. See Phosphatidylinositol bisphosphate
PIP3. See Phosphatidylinositol trisphosphate
PIXα, 305
PKA. See Protein kinase A
PKBR1, 191–192
PKC. See Protein kinase C
PKD. See Protein kinase D
PKG. See Cyclic GMP-dependent protein kinase
PKG. See Protein kinase G
PKI. See Protein kinase inhibitor
PKR. See Double-stranded RNA-dependent kinase
PLA2. See Phospholipase A2
Platelet-derived growth factor (PDGF)
chemokine activity, 193
receptor, 9
PLC. See Phospholipase C
Plk1. See Polo-like kinase, 1
Plx1, 338
Plzf, 334
PML-RAR fusion, 417
Polarity. See Cell polarity
Polo-like kinase, 1 (Plk1), 40–41, 338
Pom1, 157
Pop1, 146
Postsynaptic density (PSD)
calcium-regulated signaling
calcineurin, 253–254
calcium/calmodulin-dependent protein kinase II, 251–253, 258
scaffold proteins, 254–256
PP1. See Protein phosphatase, 1
PP2A. See Protein phosphatase, 2A
PPAR. See Peroxisome proliferator-activated receptor
PRAS40, 88
Pregnane X receptor (PXR), 23–24
Prex1, 305
Prolyl hydroxylase (PHD), 171–173
Prostaglandins, signaling overview, 59
Protease-activated receptor, 1 (PAR1), 10
Protein kinase A (PKA)
cyclic AMP target, 53–54, 101
isozymes, 57
myristoylation, 43
regulation, 101
substrates, 101
Protein kinase B. See Akt
Protein kinase C (PKC)
atypical PKC, 135, 188, 202–203, 205–208
diacylglycerol target, 55, 57
feedback control, 9
lipid messengers, 57
lymphocyte signaling, 318–319
muscle calcium sensitization, 273
Par protein localization, 205–207
polarity signaling, 207–209
receptor feedback, 9
Protein kinase D (PKD), 318
Protein kinase G (PKG), cyclic GMP target, 55
Protein kinase inhibitor (PKI), 101
Protein levels, equation, 45
Protein phosphatase, 1 (PP1)
learning and memory role, 254
PKA as substrate, 101
Protein phosphatase, 2A (PP2A)
CDK1 as substrate, 155
learning and memory role, 254
oocyte maturation role, 331
PKA as substrate, 101
PSD. See Postsynaptic density
PSD93, 254
PSD95, 254–256
P-selectin, 303
PSGL1, 303
PTB domain, protein–protein interactions, 39–40
PTEN, 59, 89, 176, 189–191, 407–408, 418
PTP. See Permeability transition pore
Puma, 372
PXR. See Pregnane X receptor
Pyk2, 18
Pyrin domain, 368
Pyruvate dehydrogenase kinase (PDK), 284, 414
Pyruvate kinase, 173–174, 415, 431
Q
Q30, 400
R
RA. See Retinoic acid
Rab, 104, 387, 431, 82, 141, 188–189, 396–398
cell migration role, 188–189
G-protein-coupled receptor signaling, 13
Rac1, 354, 368
RACK1, 350
Rad17, 159
Raf1, synaptic plasticity role, 258
Rag, 175
RAIDD, 374–375
RAMPs. See Receptor activity-modifying proteins
Rap
learning and memory role, 259
synaptic plasticity role, 258–259
Rap1, 318
RAR. See Retinoic acid receptor
Ras
cancer signaling, 408–410, 412, 415–416, 418, 421
innate immunity, 305
learning and memory role, 259
lymphocyte signaling, 319
prenylation, 42–43
synaptic plasticity role, 257–259
RasC, 192
Rb. See Retinoblastoma protein
Receptor activity-modifying proteins (RAMPs), 14
Receptor tyrosine kinases (RTKs)
cell adhesion molecule interactions, 21
coreceptors, 8
dimerization, 5–7
downstream signaling, 8
endocytosis, 9
feedback and amplification, 9
mutations and disease, 9
overview, 4–5
proteolysis, 10
Regulators of G-protein-coupled receptor signaling (RGS), 15
Replication protein A (RPA), 158
Reproduction. See also Meiosis; Oocyte; Sperm
fertilization
acrosome reaction, 336
gamete fusion and egg activation, 336–338
prospects for study, 339
zygote formation, 338
Ret, 333–334
Retinoblastoma protein (Rb)
cancer, 408
cell cycle control, 140–141
Retinoic acid (RA), sperm maturation role, 335
Retinoic acid receptor (RAR), 22, 317
Retinoid X receptor (RXR), 21–22, 130
REV-ERB, 24–25
RGS. See Regulators of G-protein-coupled receptor signaling
Rheb, 175, 322
RHIM, 298, 303, 380
Rho, G-protein-coupled receptor signaling, 13
Rho1, 207
RhoA, 396
cell migration role, 188
cytoskeleton regulation, 188
Par6 regulation, 208
Rhodopsin, 239–240
RhoG, 305
RIG-I-like receptor (RLR), signaling, 298–300
RIP1, 298, 300, 303, 355, 373–374, 380
RIP2, 300
RIP3, 303, 366, 373, 379–380
RIPK, 380
RLR. See RIG-I-like receptor
ROCK, 186
Ror, 106
RPA. See Replication protein A
RSK, 330–331, 411
Rsr1, 200
RTKs. See Receptor tyrosine kinases
Rub1, 434
Rum1, 147
RXR. See Retinoid X receptor
Ryanodine receptor (RyR), 264–265, 267–269, 271, 274
Ryk, 8, 105
RyR. See Ryanodine receptor
S
S6 kinase, 88
S144, 205
SAP102, 254
SAP97, 254
SAPK. See p38 stress-activated protein kinase
SARA. See Smad anchor for receptor activation
SCF. See Stem cell factor
Scribble, 135, 204, 210, 416
SDF1, 37
Second messengers. See also specific molecules
cyclic nucleotides, 53–55
ions, 59–83
lipids, 55–59
overview, 52–53
Secretion systems, bacteria, 391–39
Sensory receptors
evolution, 242–244
G-protein-coupled receptors, 234–235, 238–239, 241
olfaction, 242
photoreceptors, 239, 241, 243
prospects for study, 244–245
receptor activation, 238–240
signaling overview, 234–238
thermosensation, 242
Septation initiation network (SIN), 161–162
SERCA, 62, 268–270, 274
Serine/threonine kinase receptors
activation, 7
downstream signaling, 8
mutations and disease, 9
overview, 4, 6
Serpent, embryonic patterning, 219
Sevenless, 230
SF. See Sperm cytosolic factor
SH2 domain
motifs, 40–41
protein–protein interactions, 39–40
SH3 domain
motifs, 41
protein–protein interactions, 39–40
SHANK, 254–256
Shc, 8
SHIP, 323, 338
Ship1, 190
SHP1, 323
Sic1, 146, 434
SIK1, 287
Sildenafil, 55
SIN. See Septation initiation network
SIRT1, 284, 287, 375
Skeletal muscle. See Muscle contraction
Ski, 113
SKP2, 146–147
SLP76, 128, 316
SMAC. See Supramolecular activation cluster
Smac, 369
Smad, transforming growth factor-β signaling, 113–114
Smad anchor for receptor activation (SARA), 113
Smo, 107
Smooth muscle. See Muscle contraction
SNAREs, 62, 336
SNARK/NUAK2, 283
SNoN, 113
SOCE. See Store-operated calcium entry
Sodium/potassium ATPase, 59
SopE, 395–396
Sperm cytosolic factor (SF), 338
Sperm. See also Reproduction
capacitation and calcium channels, 335–336
maturation
overview, 332–334
stem cell proliferation and maintenance, 334–335
spermatocyte meiosis and release, 335
Sphingomyelin, signaling, 56
Sphingosine, hydrolysis, 59
Spindle assembly checkpoint, 160–161
SPIRE, 399
SptP, 396
Src, 19, 189, 315, 411
SREBP. See Sterol response element-binding protein
SRF, 274
STATs. See JAK/STAT signaling
Stem cell factor (SCF), 145, 147, 333
Sterol response element-binding protein (SREBP), 288, 414
STIM1, 307, 316
STIM2, 316
STING, 436
Store-operated calcium entry (SOCE), 337
Stress-activated protein kinase. See p38 stress-activated protein kinase
SuFu, 107
SUMO, 45, 131
Supramolecular activation cluster (SMAC), 318
Swe1, 157
Syk, 4, 315–316
SynGAP, 258
T
TAB2, 355
TAB3, 355
TACE. See ADAM17
TAK1, 45, 295–297, 300, 303, 355
Talin, 18, 187
TAO1, 135
Target of rapamycin. See Mammalian target of rapamycin
Taste. See Sensory receptors
TAZ, 115, 230
TBC1D1, 283
TBC1D4, 282
TBK1, 123, 298, 300
T cell
classification, 314
costimulatory molecules, 319–320
cytokine signaling, 320–321
PI3K/Akt signaling, 321–323
receptor. See T-cell receptor
T-cell receptor (TCR)
adaptor molecules, 316
ITAM, 315–316
signaling
calcium, 316–317
diacylglycerol, 316–318
ERK1/2, 319
inhibitory signals, 323
nuclear factor of activated T cells, 317–318
overview, 125–127, 317
protein kinase C, 318–319
Ras, 319
structure and function, 314–315
TCF, 227, 319
TCF/LEF, 103–104
TCR. See T-cell receptor
Tel1, 158
Tem1, 162
TET1, 407
TGF-β. See Transforming growth factor-β
Thrombospondin (Tsp1), 420
Thyroid hormone
energy homeostasis role, 279–280
receptor, 22
TIGAR, 415
TIMP3, 420
Tinman, 219
Tip60, 35
TLR. See Toll-like receptor
TNFRs. See Tumor necrosis factor receptors
Toll-like receptor (TLR)
interferon induction, 297–298
ligands, 295
signaling overview, 121–123
TAK1 and IKK activation, 295–296
TLR4 signaling, 296
TRIF in signaling, 298
TopBP1, 159
TorC2, 191–192
Torso, development role, 217
Tp12, 297
TRADD, 16, 298, 300, 302, 371, 373–374
TRAF, 15
TRAF2, 59, 350–351, 355, 374
TRAF3, 123, 297–298
TRAF6, 45, 114, 295, 297, 307
TRAIL, 373, 412
Transcriptional regulation, protein levels, 43–44
Transforming growth factor-β (TGF-β)
receptor types, 8
signaling overview, 114–115
Translational regulation, protein levels, 44
TRB3, 350
TRIF, Toll-like receptor signaling, 298
Troponin C, 62
TRPA1, 242
TRPC6, 272
TRPM6, 63
TRPM7, 63
TRPM8, 239, 242
TRPV1, sensory role, 237, 239, 242, 244
TSC1, 93, 350
TSC2, 88, 93, 350, 411, 413, 416
Tsp1. See Thrombospondin
TTP, 356
Tumor necrosis factor receptors (TNFRs)
activation, 15
caspase-8 activation in death receptor pathway, 371, 373–374
ligand diversity, 15
pathology, 17
signaling, 15–17
structure, 15–16
TNFR1 signaling
cell death induction, 303
MAPK, 301–303
nuclear factor-κB, 301–303
TXNIP, 351
U
UbcH proteins, 295, 301
Ubiquitylation
G1 regulation
CIP/KIP degradation, 147–148
cyclin degradation, 145–147
pathogen disruption, 400–401
protein degradation, 44–47
UCP. See Uncoupling protein
ULK1, 378–379
Uncoupling protein (UCP), 289
Unfolded protein response (UPR)
canonical signaling, 346–349
noncanonical aspects, 349
physiological roles
cell survival and death responses, 350
inflammation, 350–351
metabolic responses, 351–353
overview, 349–350
UPR. See Unfolded protein response
V
VAMP8, 378
Vascular cell adhesion molecule (VCAM), 17, 303
Vascular endothelial growth factor (VEGF)
cancer angiogenesis, 420
hypoxia signaling, 24
VCAM. See Vascular cell adhesion molecule
VEGF. See Vascular endothelial growth factor
VHL, 171–172, 420
Vinculin, 18
VirG, 399
Virus. See Infection
Vision. See Sensory receptors
VOCC. See Voltage-operated calcium channel
Voltage-operated calcium channel (VOCC), 59, 62, 337
VopA/P, 394
VopS, 397
VPA0450, 399
Vps15, 378
Vps34, 378
W
WASP, 398
WAVE, 186, 40
Wee1, 155, 157, 329–330
WH2 domain, 399–400
WIP1, 355
Wnt
canonical signaling, 104
cell polarity signaling cross talk, 209
embryonic patterning, 225
noncanonical signaling, 104–105
signaling overview, 103–105
WTS, 134
X
XBP1, 347, 349–351, 353
XIAP, 369, 375, 412
Y
YAP, 115, 230
YK1. See Yorkie
YopE, 396
YopH, 392
YopJ, 393–394
YopT, 396
Yorkie (YKI), 134, 143, 298, 230
YpkA, 396
Z
Zap70, 4
ZO1, 20
ZO2, 20