Вы находитесь на странице: 1из 10

WiskottAldrich Syndrome, Leukocyte Adhesion Deficiency, and Other

Migration Defects in Human Primary Immunodeficiency


Dale A Moulding and Adrian J Thrasher, UCL Institute of Child Health, London, UK
2016 Elsevier Ltd. All rights reserved.

Glossary
Chemotaxis Directional migration toward a biochemical Microthrombocytopenia Platelets are unusually small and
stimulus. present in lower than normal numbers.
Hypogammaglobulinemia Deciency of all types of Myelokathexis Neutrophil retention in the bone marrow.
immunoglobulins.

Abstract

Migration is essential for immune system function, from the establishment of the hematopoietic system in the bone
marrow through to the capture and killing of invading pathogens. This reliance on migration is demonstrated by the existence
of a subset of inherited immunodeciencies that are a result of mutations that impact on migratory processes. Much of our
understanding of the importance of migration in the immune response comes from the study of these immunodeciencies.
WiskottAldrich syndrome is the rst described and best studied migratory immunodeciency, and this article will focus on
this disorder, and its wide ranging impact on immune function. We will also give an overview of other migration defects that
lead to immunodeciency, including their genetic, biochemical, and cell biological basis.

Introduction (see Table 1 for clinical signs and treatment options). The
condition was rst described in 1937 by Dr Alfred Wiskott
The immune system serves to protect the host from invading (Wiskott, 1937), and again in 1954 by Dr Robert Anderson
pathogens and malignancies by the recognition of foreign mole- Aldrich (Aldrich et al., 1954), and has been well documented
cules (antigens) that are not naturally present in the host. in many recent reviews (Moulding et al., 2013; Bouma et al.,
Immune cells then act to destroy these pathogens, infected or 2009; Thrasher and Burns, 2010; Massaad et al., 2013). The
malignant host cells expressing the antigenic molecules. This is initial description by Alfred Wiskott was for three brothers,
achieved by constant surveillance of the host for the presence and Robert Aldrich subsequently described the condition
of antigens, through the passive migration of immune cells in only affecting males from six generations of a single family,
circulating blood and lymph systems, and active migration strongly suggesting an X-linked inheritance. This X-linked
through tissues and organs (Forster and Sozzani, 2013). Migra- inheritance was conrmed in 1994 when the gene responsible
tion is key to the success of the immune system, not just in being was identied, located on the X-chromosome (Derry et al.,
able to nd antigens, but also at almost every stage of an 1994). The WAS gene encodes for a 502 amino acid protein
immune response. This dependence on migration is exemplied that plays an important role in the control of the actin cytoskel-
by the primary immune deciency (PID) known as Wiskott eton. There are over 300 disease-causing mutations described
Aldrich syndrome (WAS). Additionally, a variety of other migra- in the WAS gene (Massaad et al., 2013; Jin et al., 2004; Ochs
tory immunodeciencies exist, further highlighting the impor- and Thrasher, 2006; Albert et al., 2010), that give rise to three
tance of migration in the immune response. The study of these distinct conditions: the disease for which the gene is named,
PIDs has helped reveal the numerous migratory steps involved WAS; a milder form of the disease called X-linked thrombocy-
in a successful immune response, and the molecular and cellular topenia (XLT); and a disease with a surprisingly different
processes that are involved. These include the ability to attach to molecular basis called X-linked neutropenia (XLN).
blood vessels in order to migrate into tissues, defective signaling
from cell surface receptors, and problems with controlling the
Actin Cytoskeletal Restructuring by WASp
cytoskeleton to change cell shape and generate the force neces-
sary for directed migration. The actin cytoskeleton is the driving force behind cellular
migration (Salbreux et al., 2012; Pollard and Cooper, 2009),
so the dysfunctional control of the cytoskeleton in WAS
WiskottAldrich Syndrome patients leaves immune cells unable to migrate efciently.
WiskottAldrich syndrome protein (WASp) is the founding
Symptoms, Discovery, and Genetics
member of a family of proteins that control the formation
WAS is characterized by a classical triad of symptoms: recurrent of branched (dendritic) actin laments through control of
bacterial infections, eczema, and microthrombocytopenia a catalytic complex of proteins termed the ARP2/3 complex

416 Encyclopedia of Immunobiology, Volume 5 http://dx.doi.org/10.1016/B978-0-12-374279-7.18010-5


Immune Deficiency j WiskottAldrich Syndrome, Leukocyte Adhesion Deficiency, and Other Migration Defects 417

Table 1 Clinical signs and treatment options for migratory immunodeciencies

Disease Clinical signs Therapeutic options References

WiskottAldrich syndrome (WAS) Eczema, thrombocytopenia, prone (for classical severe WAS) Ochs et al. (2009), Aiuti et al.
(and WAS Interacting Protein to recurrent bacterial, fungal, and Prophylactic immunoglobulin and (2013), and Hacein-Bey Abina
(WIP) deciency) viral infections, bloody diarrhea, antibiotics. Denitive treatment et al. (2015)
increased tendency to bleed with hematopoietic stem cell
leading to petechiae or bruising transplant (HSCT) or possibly
gene therapy (under evaluation)
X-linked neutropenia Recurrent bacterial infections, G-CSF therapy and prophylactic Devriendt et al. (2001), Ancliff
neutropenia, monocytopenia, antibiotics according to clinical (2003), Ancliff et al. (2006), and
possible myelodysplasia severity Beel et al. (2009)
Dock8 deciency Severe food or environmental Prophylactic immunoglobulin, Engelhardt et al. (2009) and Zhang
allergies, otitis media, antibiotics, and antiviral agents. et al. (2009), Su, 2010, Chu et al.
pneumonia or bronchitis, Denitive therapy HSCT (2012), and Aydin et al. (2015)
Staphylococcal skin infections,
eczema, eosinophilia, IgE
dysregulation, and severe
cutaneous viral infections
Coronin 1A mutation EBV-induced B cell Denitive therapy HSCT Moshous et al. (2013), Punwani
lymphoproliferative syndrome, et al. (2015)
recurrent bacterial and fungal
infections, particularly ear, nose,
and throat. T cell lymphopenia
Rac2 neutrophil immunodeciency Severe recurrent bacterial Prophylactic antibiotics and Ambruso et al. (2000)
syndrome infections, poor wound healing antifungal agents. Denitive
therapy HSCT
RhoH mutation in Disseminated at warts with HPV3, Topical antiviral agents, laser Crequer et al. (2012) and Jablonska
epidermodysplasia verruciformis at red or red-brownish plaques, therapy. Possible denitive et al. (1979)
and depigmented pityriasis therapy HSCT, but may be
versicolor-like lesions with HPV 4 residual keratinocyte defects
Neutrophil actin dysfunction Severe recurrent bacterial infection. Denitive therapy HSCT Camitta et al. (1977) and Coates
syndrome Neutrophil phagocytosis and et al. (1991)
chemotaxis defective
Warts, hypogammaglobulinemia, Recurrent bacterial infections, Topical antiviral agents, laser Wetzler et al. (1990) and
infections and myelokathexis susceptibility to HPV and therapy. Prophylactic McDermott et al. (2011)
syndrome development of warts immunoglobulin infusions,
G-CSF and GM-CSF injections.
CXCR4 antagonist therapy in trial
Leukocyte adhesion deciency Recurrent bacterial infections, Denitive therapy HSCT Elhasid and Rowe (2010)
neutrophilia

(a) (b)
(Welch et al., 1997; Mullins et al., 1998; Goley and Welch, EVH1 B GBD Branched actin network
PPP

n
tio
er in
iza

2006). WASp has no intrinsic catalytic activity, but serves as VCA


lym Act

Podosomes
a scaffold and activator for the ARP2/3 complex. Through inte-
po

gration of a wide variety of signals, WASp activates the ARP2/3 WASp activation
ARP2/3
complex to direct polymerization of new, branched actin la-
A

ments (Figure 1). This activation is subject to tight temporal


VC

Cdc42
and spatial regulation, meaning that the actin cytoskeleton is
P
PP

EVH1 B GBD
modied in strictly dened regions of the cell, allowing for Migration
the highly localized restructuring needed for processes such
as migration and phagocytosis. Figure 1 WiskottAldrich syndrome protein (WASp) activation and
WASp is one of the family of eight proteins that control the function during immune cell migration. (a) WASp domain structure
ARP2/3 complex, but is the only ARP2/3 activator whose showing the inactive form with the VCA domain sequestered by the
GTPase-binding domain. Activation by GTP-bound Cdc42 (and other
expression is restricted to the immune system (Derry et al.,
activating molecules see main text) releases the VCA domain allowing
1994). The other members of the WASp family are more widely
binding and activation of the ARP2/3 complex. This in turn catalyzes
expressed in multiple cell lineages (Suetsugu et al., 1999; Miki the polymerization of a new branch of F-actin from a preexisting actin
et al., 1996; Campellone et al., 2008; Zuchero et al., 2009; ber. (b) WASp is required to produce podosomes and the dendritic
Linardopoulou et al., 2007). All WASp family proteins perform branched actin network in the lamellipodia of migrating cells.
the same function: they integrate signals to control activation of Podosomes provide adhesion and traction, while the growing dendritic
the ARP2/3 complex. It is signicant that the immune system actin network provides force. Together they drive migration.
418 Immune Deficiency j WiskottAldrich Syndrome, Leukocyte Adhesion Deficiency, and Other Migration Defects

has evolved to express a unique ARP2/3 regulator, further indi- liver to the bone marrow (Greer et al., 1989; Lacout et al.,
cation of the importance of cytoskeletal restructuring and 2003). Because female cells have two copies of the
migration in these cells. X-chromosome, one of which is always inactivated, it is
WASp activity is regulated by signaling from the cell possible to test if there is preferential inactivation of the X-chro-
surface in response to a wide variety of stimuli, including che- mosome containing the defective WAS gene. Initially during
moattractants, integrin binding, cytokines, direct cellcell development of the hematopoietic system in mice, there is
interactions, and even the local curvature of the cell a random inactivation of the X-chromosome in fetal hemato-
membrane (Padrick et al., 2008; Gallop et al., 2013). WASp poietic cells, with some cells expressing wild-type WAS and
is a multidomain protein containing ve distinct domains: some expressing mutant WAS. However, after migration
the N-terminal EVH1 domain, the basic domain, the GTPase of the hematopoietic system to the bone marrow,
binding domain (GBD), the polyproline domain, and the X-chromosome inactivation is nonrandom, favoring inactiva-
VCA domain located at the C-terminal (Blundell et al., tion of the defective WAS X-chromosome. This demonstrates
2010). Before activation, WASp exists in the cytosol in an a distinct homing advantage in WAS normal hematopoietic
autoinhibited conformation, with the VCA domain tethered progenitors. However, as WAS patients and mice have normal
to the central GBD domain (Figure 1). Activation is mediated numbers of hematopoietic bone marrow cells, the migration
by binding of GTP-bound Cdc42 to the GBD domain, phos- deciency is not absolute. Cells expressing normal WASp
phorylation of Tyr291 also in the GBD, and binding of phos- have a migration and homing advantage, so they establish in
phatidylinositol 4,5-bisphosphate to the basic domain. the bone marrow before the WASp mutants. In the absence
Additionally, there is an association with an expanding list of competition from WASp procient cells, the WASp mutants
of kinases, adapter proteins, and actin-binding proteins that still establish a full hematopoietic system in the bone marrow.
associate with the polyproline domain (Campellone and
Welch, 2010; Thrasher and Burns, 2010; Blundell et al., Phagocytic Cell Migration
2010). Optimal activation of WASp may require coordinated The formation of branched (dendritic) actin networks through
input from a variety of these activators, with this multilevel the regulated activation of WASp is central to a variety of
regulation of WASp, potentially allowing for ne-tuning of immune functions. In phagocytic cells, these dendritic
its activation status. Following activation, the VCA domain networks are used to migrate toward, capture and phagocytose
is released from its inhibitory association with the GBD and invading pathogens, and also in the immune synapse formed
can bind to and activate the ARP2/3 complex, which in turn during intercellular interactions between phagocytes and
catalyzes the formation of branched actin laments lymphocytes. Macrophages and neutrophils use the force
(Figure 1). generated by a growing dendritic actin network to form lamel-
The loss of WASp function is the underlying cause of all the lipodia and drive the cell forward during migration (Krause
cellular and immunological defects in WAS and XLT. In WAS, and Gautreau, 2014; Figure 1(b)). Additionally, small
this loss of activity tends to be more severe than in XLT, due to dense actin structures called podosomes are formed by
mutations that lead to a complete or near complete absence of WASp-mediated ARP2/3 activation in dendritic cells (DCs)
WASp expression. In XLT, the phenotype is generally milder as and macrophages (Olivier et al., 2006). These podosomes are
the mutations lead to a partial loss of WASp activity, although surrounded by vinculin and other adhesion molecules, and
the precise link between residual WASp activity and disease facilitate adhesion of migrating cells to surrounding tissue
severity is not clearly dened (Jin et al., 2004; Albert et al., (Burns et al., 2004). The loss of WASp activity severely impedes
2010; Thrasher and Burns, 2010). Indeed, exceptions to the rela- the migration of these phagocytic cells, with poor-chemotactic
tionship between disease severity and residual WASp expression response to a variety of chemoattractants including bacterial
do exist. For example, mutations that specically block WASp peptides and host-produced chemokines and cytokines (Ochs
activation of the ARP2/3 complex (Worth, 2013. Unpublished et al., 1980; Badolato et al., 1998; Cvejic et al., 2008; Snapper
data, personal communication.) result in normal WASp expres- et al., 2005; Zhang et al., 2006; Zicha et al., 1998).
sion levels, but severe disease phenotype. XLN is the third The poor-chemotactic response is due to the inability to
disease attributed to mutations in the WAS gene, but rather restructure the actin cytoskeleton in response to signals from
than the loss of WASp activity seen in WAS and XLT, there is these chemoattractants. The lack of podosome formation
an activation of WASp with disruption of the normal autoinhi- (Jones et al., 2002; Linder et al., 1999; Tsuboi, 2007; Zhang
bited conrmation, leading to excessive activity and uncon- et al., 2006) prevents the vinculin-mediated integrin clustering
trolled actin polymerization (Thrasher and Burns, 2010). required for efcient adhesion to surrounding tissues (Burns
et al., 2004; Tsuboi and Meerloo, 2007), meaning the cells
have reduced traction to migrate. Combined with the inef-
Migration Defects in WAS
cient formation of dendritic actin networks (Ishihara et al.,
Establishment of the Hematopoietic System in the Bone Marrow 2012), the lack of traction and force generation due to the
The hematopoietic system initially develops in the fetal liver, absence of WASp activity leaves phagocytic cells poorly equip-
with WASp expressed even at the very earliest stages of hemato- ped to migrate toward and capture pathogens.
poiesis (Parolini et al., 1997). Before birth, the entire hemato-
poietic system transfers to the bone marrow (Ciriza et al., DC Migration
2013). Evidence from female carriers of WAS, and female DCs are phagocytic cells of the innate immune system that
WAS mice, suggests that WASp is required for the efcient form the link between innate and adaptive immunity. They
migration of the developing hematopoietic system from the phagocytose and process antigens in order to present these
Immune Deficiency j WiskottAldrich Syndrome, Leukocyte Adhesion Deficiency, and Other Migration Defects 419

antigens to lymphocytes of the adaptive immune system. This immune cells is inefcient. Recruitment of professional antigen
whole process is heavily reliant on migration. DCs must presenting cells at inammatory sites and then delivery of
migrate to a site of infection, phagocytose, and process antigen, antigen to the adaptive arm of the immune system are similarly
then migrate to lymph nodes and spleen to present antigen to perturbed. This is then further compounded by the poor migra-
lymphocytes. The migration defects present in macrophages tion of activated adaptive immune cells such as lymphocytes
and neutrophils also apply to DCs. Failure to produce stable and natural killer cells. It is the cumulative effect of poor migra-
lamellipodia and the absence of podosomes to direct integrin tion across all hematopoietic lineages that combine to produce
clustering results in poor adhesion to intercellular adhesion the WAS phenotype.
molecule-1 on the endothelium (Binks et al., 1998; Bouma
et al., 2007; Burns et al., 2004, 2001; de Noronha et al.,
2005; Olivier et al., 2006), and reduced traction. Even if DCs Other Actin CytoskeletalRelated Migratory Defects
do succeed in migrating to inammatory sites, they subse-
X-Linked Neutropenia
quently fail to correctly localize in the lymph nodes and spleen
(Bouma et al., 2007; de Noronha et al., 2005). Together these In 2001, a disorder was described that was the result of a novel
migratory defects lead to an extremely inefcient activation of mutation in the WAS gene that resulted in an L270P mutation in
the adaptive immune system in WAS patients. The eczema the GBD of WASp (Devriendt et al., 2001). Patients with this
that is characteristic of WAS may also result from poor DC WAS mutation suffer from severe recurrent bacterial infections
migration, as the slowly migrating DCs may mature before due to a profound neutropenia (Table 1). There are now four
reaching the lymph nodes, triggering inammation in other- naturally occurring mutations all clustered within the GBD of
wise healthy tissue. WASp that give rise to the same disease (Ancliff et al., 2006;
Beel et al., 2009; Thrasher and Burns, 2010). These mutations
Lymphocyte Migration produce a WASp that has lost its autoinhibited conformation,
The hallmark defect in WAS is the abnormal formation of the and activates the ARP2/3 complex indiscriminately without
immune synapse between antigen-presenting cells and the need for any activation signal. This leads to excessive ectopic
lymphocytes (Gallego et al., 1997; Molina et al., 1993; Snapper F-actin resulting in neutropenia due to massive apoptosis of
et al., 1998; Zhang et al., 1999). However, migratory de- maturing neutrophils (Ancliff et al., 2006; Moulding et al.,
ciencies in both these cell types also contribute to the WAS 2007).
phenotype. WASp-decient T cells show poor adhesion (so The presence of excess F-actin alters the mechanical proper-
lack traction) and fail to migrate in response to chemokines ties of the cell (Moulding et al., 2007, 2012), and this in turn
such as CCL19 and CCL21 (Snapper et al., 2005). Even after has an impact on migration. In model systems where constitu-
the suboptimal activation and proliferation of T cells by DCs, tively active WASp is expressed in myeloid cell lines, migration
the few resulting activated T cells have poor homing to the in response to the bacterial chemoattractant fMLP is far slower
inammation site (Bouma et al., 2009). Similarly B lympho- than controls (Record et al., 2013. Unpublished observations.).
cytes adhere poorly, fail to polarize the cytoskeleton, and Surprisingly, despite the slow migration, the directionality of
migrate inefciently toward the chemokine CXCL13 and sphin- these cells toward the chemoattractant is improved. It is likely
gosine-1-phosphate (Westerberg et al., 2001, 2005; Ansel et al., that the slow migration and improved directionality are attribut-
2000; Meyer-Bahlburg et al., 2008; Park et al., 2005). able to the decreased elasticity of these cells, where the inability
In addition to the role of podosomes in adhesion to to rapidly restructure the actin cytoskeleton results in slow
surrounding tissue, studies in lymphocytes revealed a novel migration and fewer direction changes.
role for podosomes during migration through the endothe-
lium. Lymphocytes extend invasive protrusions with a podo-
DOCK8 Deciency
some at their core that probe the surface of the endothelium
to nd sites permissive to transmigration. They then extend In 2009, a new severe combined immunodeciency (SCID)
the podosome directly through an endothelial cell forming was described with the discovery of mutations in the
a pore through which the lymphocyte migrates across the endo- DOCK8 gene in patients previously diagnosed with an auto-
thelium. This process is dependent on WASp (Carman et al., somal recessive form of hyper IgE syndrome (Zhang et al.,
2007), and is very likely to also play a role in transendothelial 2009; Engelhardt et al., 2009). Unlike WAS, which is expressed
migration in a variety of immune cells (Carman and Springer, solely in the immune system, DOCK8 is expressed in most
2008). tissues, but is most abundant in hematopoietic cells (Ruusala
and Aspenstrom, 2004). DOCK8 regulates the actin cytoskel-
Summary eton through its role as a guanine exchange factor, converting
The basis for the diverse symptoms seen in WAS and XLT is GDP to GTP to activate Rho GTPases. One of the Rho GTPases
a complex mix of defects in almost all immune cell lineages, activated by DOCK8 is Cdc42, which in turn activates WASp
many of which may be directly related to migration. The migra- (and other cytoskeletal regulators) (Miyamoto and Yamauchi,
tory defects are not absolute: WASp-decient cells are able to 2010). As a key regulator of the actin cytoskeleton, the absence
migrate, but do so with low speed and poor directionality. of DOCK8 causes a broad range of immune cell defects.
Slow migration of innate immune cells to sites of infection There are relatively few reports of specic migration defects
allows more time for infections to develop. The lack of inltra- due to DOCK8 deciency, but where migration has been
tion of innate immune cells means there are few cells present to assessed, it is similar to the defects seen in WAS. DCs show
release inammatory cytokines, so recruitment of further cytoskeletal reorganization defects and slow migration through
420 Immune Deficiency j WiskottAldrich Syndrome, Leukocyte Adhesion Deficiency, and Other Migration Defects

tissue to infections and lymph nodes (Harada et al., 2012). The with defects in upstream regulators such as DOCK8 and down-
cytoskeletal defects in WAS are also mirrored in DOCK8-de- stream effectors such as WASp leading to SCID. The Rho GTPases
cient natural killer cells, with failure to polarize the actin cyto- sit in the middle of the pathway, transducing signals from the
skeleton and a lack of integrin clustering (Mizesko et al., 2013). upstream regulators to the downstream effectors, so it is not
While many of the cellular defects in DOCK8 deciency are surprising that PIDs can result from mutations in Rho GTPase
shared with WAS, the SCID in DOCK8 deciency has a far genes. In 2000, two patients were independently found to
broader clinical impact (Table 1), presumably as DOCK8 have a PID-termed neutrophil immunodeciency syndrome
controls many cytoskeletal regulators in addition to WASp with mutations in the RAC2 gene (Ambruso et al., 2000;
including N-WASp. Williams et al., 2000). Rac2 is only expressed in the immune
lineage, and is particularly abundant in neutrophils. A third
case was described in 2011 (Accetta et al., 2011), with all three
Exceptionally Rare Actin Cytoskeletal Immunodeciencies
cases sharing an identical D57N mutation resulting in a domi-
WIP Deficiency nant negative protein. Neutrophils expressing dominant nega-
WASp requires the chaperone WIP to stabilize the protein and tive Rac2 are unable to restructure and polarize the actin
protect it from proteolysis (Chou et al., 2006; de la Fuente cytoskeleton so show very poor chemotaxis. Thymic T cell
et al., 2007; Konno et al., 2007). Therefore, loss of WIP expres- production is also perturbed, a phenotype shared with Coronin
sion may give rise to a WASp deciency. In 2012, a report 1A SCID, possibly due to defective integrin function. The
described a female patient with all the symptoms of WAS extreme rarity of this disorder has limited patient studies.
(Lanzi et al., 2012). The patient had normal WAS gene However, studies in mice have shown that Rac2 mutations
sequence and mRNA, but lacked WASp expression. Further have a broad impact on many immune cell functions, with
investigation revealed a mutation in the WIP gene resulting migratory and other defects in hematopoietic stem cells
in a truncated protein that was unable to bind to and stabilize (HSCs) (Gu et al., 2003), B and T lymphocytes (Croker et al.,
WASp. This novel immunodeciency most likely has all the 2002; Li et al., 2000; Walmsley et al., 2003), macrophages
migratory defects present in WAS, but the additional loss of (Yamauchi et al., 2004), and eosinophils (Fulkerson et al.,
WIP may lead to further complications. Recent reports demon- 2005).
strate that independent of its function stabilizing WASp, WIP In 2012, a loss of function mutation in the RhoH gene was
also binds actin and stabilizes the cytoskeleton to aid T cell identied as a novel cause of epidermodysplasia verruciformis
migration, adhesion, and homing (Massaad et al., 2014). (Crequer et al., 2012). This disease is characterized by persis-
tent EV-human papilloma virus infections, and T cell homing
Coronin 1A Mutations in these two patients is decient, due to lack of b7 integrin
In 2008, a spontaneously occurring disease in mice, peripheral T expression. The cellular basis of this disorder is probably a result
celldecient mice, and a novel SCID were shown to be a result of excessive actin production, as RhoH inhibition of the actin
of a Coronin 1A mutation (Shiow et al., 2008). The major defect polymerizing signaling from Rac2 is lost (Chae et al., 2008;
in Coronin 1Aassociated SCID is an absence of peripheral T Gu et al., 2005).
cells, despite normal thymic development. This has been These two Rho GTPase disorders result in opposing cyto-
explained as a migration defect, where T cells are unable to skeletal defects, loss of F-actin production in Rac2 deciency,
exit the thymus. Since the initial report, a total of seven cases and excessive actin production in RhoH deciency. This high-
of Coronin 1A SCID have been identied, all sharing the same lights the need for delicate regulation of the actin cytoskeleton
thymic egress defect (Punwani et al., 2015). Coronin 1A is for efcient immune cell function.
involved in actin cytoskeletal remodeling, but its exact mode
of action remains unclear (Chan et al., 2011). It can bind the Neutrophil Actin Dysfunction Syndrome
plasma membrane, F-actin, the ARP2/3 complex, other actin In 1991, an immunodeciency affecting three siblings was
regulators such as colin and actin inhibitory protein 1, and described with severe neutrophil migration defects (Coates
signaling molecules such as PLC (Punwani et al., 2015). Coronin et al., 1991). While the genetic basis for the disease remains
1A can promote actin disassembly by holding the ARP2/3 unknown, two proteins are overexpressed, a 47 kDa protein
complex in an inactive formation (Machesky et al., 1997), or later identied as LSP1, and an unknown 87 kDa protein
triggering colin-mediated actin depolymerization (Kueh et al., (Howard et al., 1994). LSP1 is an actin bundling protein that
2008). The lack of Coronin 1A activity leads to increased organizes F-actin into bundles in lamellipodia, lopodia, and
F-actin content, and this was initially thought to be the cause membrane rufes (Jongstra-Bilen and Jongstra, 2006). Loss of
of the failure of thymic egress (Fger et al., 2006). However, LSP1 expression enhances chemotaxis in mouse models
the signaling through PLC may be more important, with a lack (Jongstra-Bilen et al., 2000), suggesting that the overexpression
of calcium signaling and cytoskeletal restructuring leading to T of LSP1 in neutrophil actin dysfunction syndrome may impede
cell migration defects and apoptosis (Mueller et al., 2011, neutrophil migration through excessive actin bundling and an
2008). Coronin 1A deciency highlights the need for migration inability to restructure the cytoskeleton.
not just in getting to and from sites of infection, but also in the
successful development of functional immune cells.
Signaling and Adhesion Defects
Rho GTPase Mutations
The Rho GTPase pathway is the main regulator of actin cytoskel- All of the migratory SCIDs described above involve defects in
etal dynamics (Heasman and Ridley, 2008; Mulloy et al., 2010), the control of actin cytoskeletal restructuring. Migration is
Immune Deficiency j WiskottAldrich Syndrome, Leukocyte Adhesion Deficiency, and Other Migration Defects 421

driven by cytoskeletal restructuring, but is triggered in response of fucose modication on certain cell surface glycoproteins
to external stimuli and signaling, and requires adhesion on leukocytes and endothelial cells that serve as ligand for
between the migrating cell and its environment. Defects the selectin family of surface receptors (van de Vijver et al.,
in both of these processes also lead to migration 2013). Selectins are important for the initial attachment
immunodeciencies. and rolling of leukocytes along the endothelial cells of the
vessel wall. Immunologically, LAD II is a less-severe disease
than LAD I, as some extravasation still occurs; however,
Warts, Hypogammaglobulinemia, Infections,
other severe developmental and mental defects are also
and Myelokathexis
present due to the wider role of fucosylation outside of
Warts, hypogammaglobulinemia, infections, and myeloka- the immune system.
thexis (WHIM) was rst described in 1990 and is characterized LAD III was rst described in 1997 (Kuijpers et al., 1997), and
by recurrent bacterial infections and the development of is now known to be caused by mutations in the FERMT3 gene
a susceptibility to HPV in the teenage years leading to warts encoding Kindlin-3. This protein is involved in inside-out
on the hands and feet, and neutropenia and monocytopenia signaling that is important in integrin activation and the forma-
(Wetzler et al., 1990; Al Ustwani et al., 2014). WHIM is caused tion of high-afnity attachment to their ligands (van de Vijver
by mutations in the chemokine receptor CXCR4, generating an et al., 2013). LAD III is a severe disorder, with bleeding compli-
overactive receptor. While CXCR4 is expressed in many cells cations due to loss of integrin activation on platelets, and has
and tissues, mutations manifest primarily as an immunode- a high-mortality rate.
ciency, perhaps indicative of the particular dependence of the
immune system on migration.
CXCR4 is the receptor for the chemokine CXCL12, which Concluding Remarks
is strongly expressed on stromal cells in the bone marrow
and acts to retain HSCs. The overactivation of the CXCR4 The PIDs presented here together show how defects in migra-
receptor on neutrophils (and monocytes) inhibits the migra- tion affect almost every aspect of the immune response. Migra-
tion of these cells out of the bone marrow, leading to neutro- tion is required to establish the hematopoietic system in the
penia and monocytopenia. T cell and B cell development are bone marrow, and the resulting mature hematopoietic cells
also perturbed, although this is a result of defective migra- must then be able to leave the bone marrow and travel
tion within the thymus, spleen, and lymph nodes, rather throughout the host tissues. T cell development in the thymus
than entrapment within the bone marrow. In this syndrome, depends on correct thymic localization of developing cells, and
the migration machinery is intact, but the faulty signaling then migration drives the egress of these cells from the thymus.
results in inappropriate migratory signals and immunode- Lymphocyte activation in the spleen and lymph nodes is
ciency. Recently, one patient with WHIM syndrome was impeded by poor-migratory capacity, and homing of activated
cured by a spontaneous chromothripsis (McDermott et al., lymphocytes to inammatory sites is dependent on active
2015). Here, deletion of the mutant allele allowed selective migration. Attachment to and migration through the endothe-
repopulation of the myeloid compartment and resolution lium to leave the bloodstream and migrate to an infection is
of symptoms. often defective in migratory immunodeciencies, and the
subsequent journey through tissue to the infection site is also
perturbed.
Leukocyte Adhesion Deciency Syndrome
Any one of these migratory deciencies is sufcient to cause
In order for phagocytes to reach a site of infection, they must disease, as the immune response is an intricate and highly
rst attach to and migrate through the endothelium of the interdependent process. Failure at any level has knock on
blood vessel wall. A group of related defects called leukocyte effects, and multilevel failure as seen in WAS, combine to
adhesion deciency (LAD) form a subset of PIDs where produce severe defects, resulting in immunodeciency, autoim-
adhesion to the endothelium is disrupted, preventing extrav- munity, and predisposition to cancer. The wide ranging clinical
asation from the bloodstream into the surrounding tissues impact of these migration immunodeciencies limits treatment
(van de Vijver et al., 2013). The rst type of LAD was options, with the only denitive treatment often being HSC
described in 1980 (Crowley et al., 1980), as a disorder char- transplantation.
acterized by recurrent bacterial infections and neutrophilia
in which neutrophils are unable to adhere and spread. The
gene was subsequently identied as ITGB2, the common b-
See also: Anatomy and Microanatomy of the Immune System:
chain of the leukocyte integrin CD18, with at least 86
Blood Vascular Endothelial Adhesion Molecules; Effector T
different point mutations described in LAD I patients (van
Lymphocyte Migration to and Within Non-Lymphoid Tissues;
de Vijver et al., 2012). CD18 is required for the formation
Endothelial Adhesion Molecules in the Lymphatic Vasculature;
of a rm adhesion to the endothelium, and lack of expres-
Leukocyte Adhesion Molecules; Lymph Node Structure; Roles
of Chemokines in Immune Cell Trafcking to Lymphoid
sion severely impairs endothelial transmigration (Fiorini
Tissues; S1P and LPA: Regulators of Immune Cell Egress and
et al., 2002).
Ingress in Lymphoid Tissues; Structure and Function of the
LAD II is an extremely rare disorder rst described in
Bone Marrow Hematopoietic Niche; The Spleen; Thymic
1992 (Etzioni et al., 1992), in which the SLC35C1 gene
Microenvironments: Development, Organization, and Function.
encoding for a Golgi GDP-fucose transport protein is
Cells of the Innate Immune System: Dendritic Cells and
mutated resulting in loss of expression. This leads to a lack
422 Immune Deficiency j WiskottAldrich Syndrome, Leukocyte Adhesion Deficiency, and Other Migration Defects

Aydin, S.E., Kilic, S.S., Aytekin, C., Kumar, A., Porras, O., Kainulainen, L.,
Dendritic Cell Subsets; ILC2 in Immunity; Macrophage
Kostyuchenko, L., Genel, F., Kutukculer, N., Karaca, N., Gonzalez-Granado, L.,
Activation and Polarization; Natural Killer Cells; Neutrophils Abbott, J., Al-Zahrani, D., Rezaei, N., Baz, Z., Thiel, J., Ehl, S., Marodi, L.,
Role in Innate Immunity; The Role of Invariant NKT Cells in Orange, J.S., Sawalle-Belohradsky, J., Keles, S., Holland, S.M., Sanal, O.,
Immunity. Development of T Cells and Innate Lymphoid Cells: Ayvaz, D.C., Tezcan, I., Al-Mousa, H., Alsum, Z., Hawwari, A., Metin, A.,
CD4/CD8 Lineage Commitment; Control of Migration during Matthes-Martin, S., Honig, M., Schulz, A., Picard, C., Barlogis, V., Gennery, A.,
Ifversen, M., Van Montfrans, J., Kuijpers, T., Bredius, R., Duckers, G., Al-
Intrathymic T Cell Development; Development of Group 2 Herz, W., Pai, S.Y., Geha, R., Notheis, G., Schwarze, C.P., Tavil, B., Azik, F.,
Innate Lymphoid Cells; Development of Natural Killer Cells and Bienemann, K., Grimbacher, B., Heinz, V., Gaspar, H.B., Aydin, R., Hagl, B.,
ILC1; Development of Regulatory T Cells in the Thymus; Innate Gathmann, B., Belohradsky, B.H., Ochs, H.D., Chatila, T., Renner, E.D., Su, H.,
Lymphoid Cells Type 3; Orchestration of T Cell Development by Freeman, A.F., Engelhardt, K., Albert, M.H., On Behalf of the Inborn Errors
Working Party of EBMT, 2015. DOCK8 deciency: clinical and immunological
Common g Chain Cytokines; Production of the Semi-Invariant
phenotype and treatment options - a review of 136 patients. J. Clin. Immunol.
TCR and PLZF Function in Innate Programming of iNKT Cells; 35, 189198.
Transcriptional Regulation of T Cell Lineage Commitment; Badolato, R., Sozzani, S., Malacarne, F., Bresciani, S., Fiorini, M., Borsatti, A.,
Transcriptional and Microenvironmental Regulation of gd T Cell Albertini, A., Mantovani, A., Ugazio, A.G., Notarangelo, L.D., 1998. Monocytes
Development. Immune Deciency: Combined T Cell and B Cell from Wiskott-Aldrich patients display reduced chemotaxis and lack of cell polari-
zation in response to monocyte chemoattractant protein-1 and formyl-methionyl-
Deciency SCID Forms: TB; Primary Immunodeciencies: leucyl-phenylalanine. J. Immunol. 161, 10261033.
Clinical and Biological Guidelines. MHC Structure and Function: Beel, K., Cotter, M.M., Blatny, J., Bond, J., Lucas, G., Green, F., Vanduppen, V.,
Origin and Processing of MHC Class II Ligands; Origin and Leung, D.W., Rooney, S., Smith, O.P., Rosen, M.K., Vandenberghe, P., 2009.
Processing of MHC-I Ligands. Myeloid and B Cell A large kindred with X-linked neutropenia with an I294t mutation of the Wiskott-
Aldrich syndrome gene. Br. J. Haematol. 144, 120126.
Development: Adult Hematopoiesis; Marginal Zone B Cell
Binks, M., Jones, G.E., Brickell, P.M., Kinnon, C., Katz, D.R., Thrasher, A.J., 1998.
Development; Myelopoiesis; Ontogeny of the Hematopoietic Intrinsic dendritic cell abnormalities in Wiskott-Aldrich syndrome. Eur. J. Immunol.
System. Signal Transduction: Immunological Synapses; Signal 28, 32593267.
Transduction by the B Cell Antigen Receptor; Signaling of Blundell, M.P., Worth, A., Bouma, G., Thrasher, A.J., 2010. The Wiskott-Aldrich
Phagocytosis; TCR Signaling: Proximal Signaling. Structure syndrome: the actin cytoskeleton and immune cell function. Dis. Markers 29,
157175.
and Function of Diversifying Receptors: Organization and Bouma, G., Burns, S., Thrasher, A.J., 2007. Impaired T-cell priming in vivo resulting
Rearrangement of TCR Loci; Structure and Function of TCRab from dysfunction of WASp-decient dendritic cells. Blood 110, 42784284.
Receptors; Structure and Function of TCRgd Receptors. Bouma, G., Burns, S.O., Thrasher, A.J., 2009. Wiskott-Aldrich Syndrome: immuno-
deciency resulting from defective cell migration and impaired immunostimulatory
activation. Immunobiology 214, 778790.
Burns, S., Thrasher, A.J., Blundell, M.P., Machesky, L., Jones, G.E., 2001. Cong-
References uration of human dendritic cell cytoskeleton by Rho GTPases, the WAS protein, and
differentiation. Blood 98, 11421149.
Accetta, D., Syverson, G., Bonacci, B., Reddy, S., Bengtson, C., Surfus, J., Burns, S., Hardy, S.J., Buddle, J., Yong, K.L., Jones, G.E., Thrasher, A.J., 2004.
Harbeck, R., Huttenlocher, A., Grossman, W., Routes, J., Verbsky, J., 2011. Maturation of DC is associated with changes in motile characteristics and
Human phagocyte defect caused by a Rac2 mutation detected by means of adherence. Cell Motil. Cytoskeleton 57, 118132.
neonatal screening for T-cell lymphopenia. J. Allergy Clin. Immunol. 127, 535538 Camitta, B.M., Quesenberry, P.J., Parkman, R., Boxer, L.A., Stossel, T.P.,
e1e2. Cassady, J.R., Rappeport, J.M., Nathan, D.G., 1977. Bone marrow transplantation
Aiuti, A., Biasco, L., Scaramuzza, S., Ferrua, F., Cicalese, M.P., Baricordi, C., for an infant with neutrophil dysfunction. Exp. Hematol. 5, 109116.
Dionisio, F., Calabria, A., Giannelli, S., Castiello, M.C., Bosticardo, M., Campellone, K.G., Welch, M.D., 2010. A nucleator arms race: cellular control of actin
Evangelio, C., Assanelli, A., Casiraghi, M., Di Nunzio, S., Callegaro, L., Benati, C., assembly. Nat. Rev. Mol. Cell Biol. 11, 237251.
Rizzardi, P., Pellin, D., Di Serio, C., Schmidt, M., Von Kalle, C., Gardner, J., Campellone, K.G., Webb, N.J., Znameroski, E.A., Welch, M.D., 2008. WHAMM is an
Mehta, N., Neduva, V., Dow, D.J., Galy, A., Miniero, R., Finocchi, A., Metin, A., Arp2/3 complex activator that binds microtubules and functions in ER to Golgi
Banerjee, P.P., Orange, J.S., Galimberti, S., Valsecchi, M.G., Bif, A., Montini, E., transport. Cell 134, 148161.
Villa, A., Ciceri, F., Roncarolo, M.G., Naldini, L., 2013. Lentiviral hematopoietic Carman, C.V., Springer, T.A., 2008. Trans-cellular migration: cell-cell contacts get
stem cell gene therapy in patients with Wiskott-Aldrich syndrome. Science 341, intimate. Curr. Opin. Cell Biol. 20, 533540.
1233151. Carman, C.V., Sage, P.T., Sciuto, T.E., de la Fuente, M.A., Geha, R.S., Ochs, H.D.,
Al Ustwani, O., Kurzrock, R., Wetzler, M., 2014. Genetics on a WHIM. Br. J. Haematol. Dvorak, H.F., Dvorak, A.M., Springer, T.A., 2007. Transcellular diapedesis is
164, 1523. initiated by invasive podosomes. Immunity 26, 784797.
Albert, M.H., Notarangelo, L.D., Ochs, H.D., 2010. Clinical spectrum, pathophysi- Chae, H.D., Lee, K.E., Williams, D.A., Gu, Y., 2008. Cross-talk between Rhoh and
ology and treatment of the Wiskott-Aldrich syndrome. Curr. Opin. Hematol. Rac1 in regulation of actin cytoskeleton and chemotaxis of hematopoietic
10 (3). progenitor cells. Blood 111, 25972605.
Aldrich, R.A., Steinberg, A.G., Campbell, D.C., 1954. Pedigree demonstrating a sex- Chan, K.T., Creed, S.J., Bear, J.E., 2011. Unraveling the enigma: progress towards
linked recessive condition characterized by draining ears, eczematoid dermatitis understanding the coronin family of actin regulators. Trends Cell Biol. 21,
and bloody diarrhea. Pediatrics 13, 133139. 481488.
Ambruso, D.R., Knall, C., Abell, A.N., Panepinto, J., Kurkchubasche, A., Chou, H.C., Anton, I.M., Holt, M.R., Curcio, C., Lanzardo, S., Worth, A., Burns, S.,
Thurman, G., Gonzalez-Aller, C., Hiester, A., Deboer, M., Harbeck, R.J., Oyer, R., Thrasher, A.J., Jones, G.E., Calle, Y., 2006. WIP regulates the stability and
Johnson, G.L., Roos, D., 2000. Human neutrophil immunodeciency syndrome localization of WASP to podosomes in migrating dendritic cells. Curr. Biol. 16,
is associated with an inhibitory Rac2 mutation. Proc. Natl. Acad. Sci. U.S.A. 97, 23372344.
46544659. Chu, E.Y., Freeman, A.F., Jing, H., Cowen, E.W., Davis, J., Su, H.C., Holland, S.M.,
Ancliff, P.J., Blundell, M.P., Cory, G.O., Calle, Y., Worth, A., Kempski, H., Turner, M.L., 2012. Cutaneous manifestations of DOCK8 deciency syndrome.
Burns, S., Jones, G.E., Sinclair, J., Kinnon, C., Hann, I.M., Gale, R.E., Arch. Dermatol. 148, 7984.
Linch, D.C., Thrasher, A.J., 2006. Two novel activating mutations in the Ciriza, J., Thompson, H., Petrosian, R., Manilay, J.O., Garcia-Ojeda, M.E., 2013.
Wiskott-Aldrich syndrome protein result in congenital neutropenia. Blood 108, The migration of hematopoietic progenitors from the fetal liver to the fetal bone
21822189. marrow: lessons learned and possible clinical applications. Exp. Hematol. 41,
Ancliff, P.J., 2003. Congenital neutropenia. Blood Rev. 17, 209216. 411423.
Ansel, K.M., Ngo, V.N., Hyman, P.L., Luther, S.A., Forster, R., Sedgwick, J.D., Coates, T.D., Torkildson, J.C., Torres, M., Church, J.A., Howard, T.H., 1991. An
Browning, J.L., Lipp, M., Cyster, J.G., 2000. A chemokine-driven positive feedback inherited defect of neutrophil motility and microlamentous cytoskeleton associated
loop organizes lymphoid follicles. Nature 406, 309314. with abnormalities in 47-Kd and 89-Kd proteins. Blood 78, 13381346.
Immune Deficiency j WiskottAldrich Syndrome, Leukocyte Adhesion Deficiency, and Other Migration Defects 423

Crequer, A., Troeger, A., Patin, E., Ma, C.S., Picard, C., Pedergnana, V., Fieschi, C., Hacein-Bey Abina, S., Gaspar, H.B., Blondeau, J., Caccavelli, L., Charrier, S.,
Lim, A., Abhyankar, A., Gineau, L., Mueller-Fleckenstein, I., Schmidt, M., Taieb, A., Buckland, K., Picard, C., Six, E., Himoudi, N., Gilmour, K., Mcnicol, A.M., Hara, H.,
Krueger, J., Abel, L., Tangye, S.G., Orth, G., Williams, D.A., Casanova, J.L., Xu-Bayford, J., Rivat, C., Touzot, F., Mavilio, F., Lim, A., Treluyer, J.M., Heritier, S.,
Jouanguy, E., 2012. Human RHOH deciency causes T cell defects and Lefrere, F., Magalon, J., Pengue-Koyi, I., Honnet, G., Blanche, S., Sherman, E.A.,
susceptibility to EV-HPV infections. J. Clin. Invest. 122, 32393247. Male, F., Berry, C., Malani, N., Bushman, F.D., Fischer, A., Thrasher, A.J.,
Croker, B.A., Tarlinton, D.M., Cluse, L.A., Tuxen, A.J., Light, A., Yang, F.C., Galy, A., Cavazzana, M., 2015. Outcomes following gene therapy in patients with
Williams, D.A., Roberts, A.W., 2002. The Rac2 guanosine triphosphatase regulates severe Wiskott-Aldrich syndrome. JAMA 313, 15501563.
B lymphocyte antigen receptor responses and chemotaxis and is required for Harada, Y., Tanaka, Y., Terasawa, M., Pieczyk, M., Habiro, K., Katakai, T., Hanawa-
establishment of B-1a and marginal zone B lymphocytes. J. Immunol. 168, Suetsugu, K., Kukimoto-Niino, M., Nishizaki, T., Shirouzu, M., Duan, X., Uruno, T.,
33763386. Nishikimi, A., Sanematsu, F., Yokoyama, S., Stein, J.V., Kinashi, T., Fukui, Y.,
Crowley, C.A., Curnutte, J.T., Rosin, R.E., Andre-Schwartz, J., Gallin, J.I., 2012. DOCK8 is a Cdc42 activator critical for interstitial dendritic cell migration
Klempner, M., Snyderman, R., Southwick, F.S., Stossel, T.P., Babior, B.M., 1980. during immune responses. Blood 119, 44514461.
An inherited abnormality of neutrophil adhesion. Its genetic transmission and its Heasman, S.J., Ridley, A.J., 2008. Mammalian Rho GTPases: new insights into their
association with a missing protein. N. Engl. J. Med. 302, 11631168. functions from in vivo studies. Nat. Rev. Mol. Cell Biol. 9, 690701.
Cvejic, A., Hall, C., Bak-Maier, M., Flores, M.V., Crosier, P., Redd, M.J., Martin, P., Howard, T., Li, Y., Torres, M., Guerrero, A., Coates, T., 1994. The 47-kD protein
2008. Analysis of WASp function during the wound inammatory response increased in neutrophil actin dysfunction with 47- and 89-kD protein abnormalities
live-imaging studies in zebrash larvae. J. Cell Sci. 121, 31963206. is lymphocyte-specic protein. Blood 83, 231241.
Derry, J.M., Ochs, H.D., Francke, U., 1994. Isolation of a novel gene mutated in Ishihara, D., Dovas, A., Park, H., Isaac, B.M., Cox, D., 2012. The chemotactic defect in
Wiskott-Aldrich syndrome. Cell 78, 635644. Wiskott-Aldrich syndrome macrophages is due to the reduced persistence of
Devriendt, K., Kim, A.S., Mathijs, G., Frints, S.G., Schwartz, M., Van Den Oord, J.J., directional protrusions. PLoS One 7, e30033.
Verhoef, G.E., Boogaerts, M.A., Fryns, J.P., You, D., Rosen, M.K., Jablonska, S., Orth, G., Jarzabek-Chorzelska, M., Glinski, W., Obalek, S., Rzesa, G.,
Vandenberghe, P., 2001. Constitutively activating mutation in WASP causes Croissant, O., Favre, M., 1979. Twenty-one years of follow-up studies of familial
X-linked severe congenital neutropenia. Nat. Genet. 27, 313317. epidermodysplasia verruciformis. Dermatologica 158, 309327.
Elhasid, R., Rowe, J.M., 2010. Hematopoetic stem cell transplantation in neutrophil Jin, Y., Mazza, C., Christie, J.R., Giliani, S., Fiorini, M., Mella, P., Gandellini, F.,
disorders: severe congenital neutropenia, leukocyte adhesion deciency and Stewart, D.M., Zhu, Q., Nelson, D.L., Notarangelo, L.D., Ochs, H.D., 2004. Mutations
chronic granulomatous disease. Clin. Rev. Allergy Immunol. 38, 6167. of the Wiskott-Aldrich Syndrome Protein (WASP): hotspots, effect on transcription,
Engelhardt, K.R., Mcghee, S., Winkler, S., Sassi, A., Woellner, C., Lopez-Herrera, G., and translation and phenotype/genotype correlation. Blood 104, 40104019.
Chen, A., Kim, H.S., Lloret, M.G., Schulze, I., Ehl, S., Thiel, J., Pfeifer, D., Jones, G.E., Zicha, D., Dunn, G.A., Blundell, M., Thrasher, A., 2002. Restoration of
Veelken, H., Niehues, T., Siepermann, K., Weinspach, S., Reisli, I., Keles, S., podosomes and chemotaxis in Wiskott-Aldrich syndrome macrophages following
Genel, F., Kutukculer, N., Camcioglu, Y., Somer, A., Karakoc-Aydiner, E., Barlan, I., induced expression of WASp. Int. J. Biochem. Cell Biol. 34, 806815.
Gennery, A., Metin, A., Degerliyurt, A., Pietrogrande, M.C., Yeganeh, M., Baz, Z., Jongstra-Bilen, D.J., Jongstra, J., 2006. Leukocyte-specic protein 1 (LSP1).
Al-Tamemi, S., Klein, C., Puck, J.M., Holland, S.M., Mccabe, E.R., Grimbacher, B., Immunol. Res. 35, 6573.
Chatila, T.A., 2009. Large deletions and point mutations involving the dedicator of Jongstra-Bilen, J., Misener, V.L., Wang, C., Ginzberg, H., Auerbach, A., Joyner, A.L.,
cytokinesis 8 (DOCK8) in the autosomal-recessive form of hyper-IgE syndrome. J. Downey, G.P., Jongstra, J., 2000. LSP1 modulates leukocyte populations in
Allergy Clin. Immunol. 124, 12891302. e4. resting and inamed peritoneum. Blood 96, 18271835.
Etzioni, A., Frydman, M., Pollack, S., Avidor, I., Phillips, M.L., Paulson, J.C., Gershoni- Konno, A., Kirby, M., Anderson, S.A., Schwartzberg, P.L., Candotti, F., 2007. The
Baruch, R., 1992. Brief report: recurrent severe infections caused by a novel expression of Wiskott-Aldrich syndrome protein (WASP) is dependent on WASP-
leukocyte adhesion deciency. N. Engl. J. Med. 327, 17891792. interacting protein (WIP). Int. Immunol. 19, 185192.
Fiorini, M., Vermi, W., Facchetti, F., Moratto, D., Alessandri, G., Notarangelo, L., Krause, M., Gautreau, A., 2014. Steering cell migration: lamellipodium dynamics and
Caruso, A., Grigolato, P., Ugazio, A.G., Notarangelo, L.D., Badolato, R., 2002. the regulation of directional persistence. Nat. Rev. Mol. Cell Biol. 15, 577590.
Defective migration of monocyte-derived dendritic cells in LAD-1 immunode- Kueh, H.Y., Charras, G.T., Mitchison, T.J., Brieher, W.M., 2008. Actin disassembly by
ciency. J. Leukoc. Biol. 72, 650656. colin, coronin, and Aip1 occurs in bursts and is inhibited by barbed-end cappers.
Fger, N., Rangell, L., Danilenko, D.M., Chan, A.C., 2006. Requirement for coronin 1 J. Cell Biol. 182, 341353.
in T Lymphocyte trafcking and cellular homeostasis. Science 313, 839842. Kuijpers, T.W., Van Lier, R.A., Hamann, D., De Boer, M., Thung, L.Y., Weening, R.S.,
Forster, R., Sozzani, S., 2013. Emerging aspects of leukocyte migration. Eur. J. Verhoeven, A.J., Roos, D., 1997. Leukocyte adhesion deciency type 1 (LAD-1)/
Immunol. 43, 14041406. variant. A novel immunodeciency syndrome characterized by dysfunctional beta2
de la Fuente, M.A., Sasahara, Y., Calamito, M., Anton, I.M., Elkhal, A., Gallego, M.D., integrins. J. Clin. Invest. 100, 17251733.
Suresh, K., Siminovitch, K., Ochs, H.D., Anderson, K.C., Rosen, F.S., Geha, R.S., Lacout, C., Haddad, E., Sabri, S., Svinarchouk, F., Garcon, L., Capron, C., Foudi, A.,
Ramesh, N., 2007. WIP is a chaperone for Wiskott-Aldrich syndrome protein Mzali, R., Snapper, S.B., Louache, F., Vainchenker, W., Dumenil, D., 2003. A
(WASP). Proc. Natl. Acad. Sci. U.S.A. 104, 926931. defect in hematopoietic stem cell migration explains the nonrandom
Fulkerson, P.C., Zhu, H., Williams, D.A., Zimmermann, N., Rothenberg, M.E., 2005. X-chromosome inactivation in carriers of Wiskott-Aldrich syndrome. Blood 102,
CXCL9 inhibits eosinophil responses by a CCR3- and Rac2-dependent mechanism. 12821289.
Blood 106, 436443. Lanzi, G., Moratto, D., Vairo, D., Masneri, S., Delmonte, O., Paganini, T., Parolini, S.,
Gallego, M.D., Santamaria, M., Pena, J., Molina, I.J., 1997. Defective actin reorga- Tabellini, G., Mazza, C., Savoldi, G., Montin, D., Martino, S., Tovo, P.,
nization and polymerization of Wiskott-Aldrich T cells in response to CD3-mediated Pessach, I.M., Massaad, M.J., Ramesh, N., Porta, F., Plebani, A.,
stimulation. Blood 90, 30893097. Notarangelo, L.D., Geha, R.S., Giliani, S., 2012. A novel primary human immu-
Gallop, J.L., Walrant, A., Cantley, L.C., Kirschner, M.W., 2013. Phosphoinosi- nodeciency due to deciency in the WASP-interacting protein WIP. J. Exp. Med.
tides and membrane curvature switch the mode of actin polymerization via 209, 2934.
selective recruitment of toca-1 and Snx9. Proc. Natl. Acad. Sci. U.S.A. 110, Li, B., Yu, H., Zheng, W., Voll, R., Na, S., Roberts, A.W., Williams, D.A., Davis, R.J.,
71937198. Ghosh, S., Flavell, R.A., 2000. Role of the guanosine triphosphatase Rac2 in
Goley, E.D., Welch, M.D., 2006. The ARP2/3 complex: an actin nucleator comes of T helper 1 cell differentiation. Science 288, 22192222.
age. Nat. Rev. Mol. Cell Biol. 7, 713726. Linardopoulou, E.V., Parghi, S.S., Friedman, C., Osborn, G.E., Parkhurst, S.M.,
Greer, W.L., Kwong, P.C., Peacocke, M., Ip, P., Rubin, L.A., Siminovitch, K.A., 1989. Trask, B.J., 2007. Human subtelomeric WASH genes encode a new subclass of
X-chromosome inactivation in the Wiskott-Aldrich syndrome: a marker for detection the WASP family. PLoS Genet. 3, e237.
of the carrier state and identication of cell lineages expressing the gene defect. Linder, S., Nelson, D., Weiss, M., Aepfelbacher, M., 1999. Wiskott-Aldrich syndrome
Genomics 4, 6067. protein regulates podosomes in primary human macrophages. Proc. Natl. Acad.
Gu, Y., Filippi, M.D., Cancelas, J.A., Siefring, J.E., Williams, E.P., Jasti, A.C., Sci. U.S.A. 96, 96489653.
Harris, C.E., Lee, A.W., Prabhakar, R., Atkinson, S.J., Kwiatkowski, D.J., Machesky, L.M., Reeves, E., Wientjes, F., Mattheyse, F.J., Grogan, A., Totty, N.F.,
Williams, D.A., 2003. Hematopoietic cell regulation by Rac1 and Rac2 guanosine Burlingame, A.L., Hsuan, J.J., Segal, A.W., 1997. Mammalian actin-related protein
triphosphatases. Science 302, 445449. 2/3 complex localizes to regions of lamellipodial protrusion and is composed of
Gu, Y., Jasti, A.C., Jansen, M., Siefring, J.E., 2005. RhoH, a hematopoietic-specic evolutionarily conserved proteins. Biochem. J. 328 (Pt 1), 105112.
Rho GTPase, regulates proliferation, survival, migration, and engraftment of Massaad, M.J., Ramesh, N., Geha, R.S., 2013. Wiskott-Aldrich syndrome:
hematopoietic progenitor cells. Blood 105, 14671475. a comprehensive review. Ann. N.Y. Acad. Sci. 1285, 2643.
424 Immune Deficiency j WiskottAldrich Syndrome, Leukocyte Adhesion Deficiency, and Other Migration Defects

Massaad, M.J., Oyoshi, M.K., Kane, J., Koduru, S., Alcaide, P., Nakamura, F., Olivier, A., Jeanson-Leh, L., Bouma, G., Compagno, D., Blondeau, J., Seye, K.,
Ramesh, N., Luscinskas, F.W., Hartwig, J., Geha, R.S., 2014. Binding of WIP to Charrier, S., Burns, S., Thrasher, A.J., Danos, O., Vainchenker, W., Galy, A., 2006.
actin is essential for T cell actin cytoskeleton integrity and tissue homing. Mol. Cell. A partial down-regulation of WASP is sufcient to inhibit podosome formation in
Biol. 34, 43434354. dendritic cells. Mol. Ther. 13, 729737.
McDermott, D.H., Liu, Q., Ulrick, J., Kwatemaa, N., Anaya-Obrien, S., Penzak, S.R., Padrick, S.B., Cheng, H.C., Ismail, A.M., Panchal, S.C., Doolittle, L.K., Kim, S.,
Filho, J.O., Priel, D.A., Kelly, C., Garofalo, M., Littel, P., Marquesen, M.M., Skehan, B.M., Umetani, J., Brautigam, C.A., Leong, J.M., Rosen, M.K., 2008.
Hilligoss, D., Decastro, R., Fleisher, T.A., Kuhns, D.B., Malech, H.L., Murphy, P.M., Hierarchical regulation of WASP/WAVE proteins. Mol. Cell 32, 426438.
2011. The CXCR4 antagonist plerixafor corrects panleukopenia in patients with Park, J.Y., Shcherbina, A., Rosen, F.S., Prodeus, A.P., Remold-Odonnell, E., 2005.
WHIM syndrome. Blood 118, 49574962. Phenotypic perturbation of B cells in the Wiskott-Aldrich syndrome. Clin. Exp.
McDermott, D.H., Gao, J.L., Liu, Q., Siwicki, M., Martens, C., Jacobs, P., Velez, D., Immunol. 139, 297305.
Yim, E., Bryke, C.R., Hsu, N., Dai, Z., Marquesen, M.M., Stregevsky, E., Parolini, O., Berardelli, S., Riedl, E., Bello-Fernandez, C., Strobl, H., Majdic, O.,
Kwatemaa, N., Theobald, N., Long Priel, D.A., Pittaluga, S., Raffeld, M.A., Knapp, W., 1997. Expression of Wiskott-Aldrich syndrome protein (WASP) gene
Calvo, K.R., Maric, I., Desmond, R., Holmes, K.L., Kuhns, D.B., Balabanian, K., during hematopoietic differentiation. Blood 90, 7075.
Bachelerie, F., Porcella, S.F., Malech, H.L., Murphy, P.M., 2015. Chromothriptic Pollard, T.D., Cooper, J.A., 2009. Actin, a central player in cell shape and movement.
cure of WHIM syndrome. Cell 160, 686699. Science 326, 12081212.
Meyer-Bahlburg, A., Becker-Herman, S., Humblet-Baron, S., Khim, S., Weber, M., Punwani, D., Pelz, B., Yu, J., Arva, N.C., Schafernak, K., Kondratowicz, K.,
Bouma, G., Thrasher, A.J., Batista, F.D., Rawlings, D.J., 2008. Wiskott-Aldrich Makhija, M., Puck, J.M., 2015. Coronin-1A: immune deciency in humans and
syndrome protein deciency in B cells results in impaired peripheral homeostasis. mice. J. Clin. Immunol. 35, 100107.
Blood 112, 41584169. Record, J., Moulding, D., Charras, G., Thrasher, A., 2013. Unpublished observation.
Miki, H., Miura, K., Takenawa, T., 1996. N-WASP, a novel actin-depolymerizing Ruusala, A., Aspenstrom, P., 2004. Isolation and characterisation of DOCK8,
protein, regulates the cortical cytoskeletal rearrangement in a Pip2-dependent a member of the DOCK180-related regulators of cell morphology. FEBS Lett. 572,
manner downstream of tyrosine kinases. EMBO J. 15, 53265335. 159166.
Miyamoto, Y., Yamauchi, J., 2010. Cellular signaling of Dock family proteins in neural Salbreux, G., Charras, G., Paluch, E., 2012. Actin cortex mechanics and cellular
function. Cell Signal. 22, 175182. morphogenesis. Trends Cell Biol. 22, 536545.
Mizesko, M.C., Banerjee, P.P., Monaco-Shawver, L., Mace, E.M., Bernal, W.E., Shiow, L.R., Roadcap, D.W., Paris, K., Watson, S.R., Grigorova, I.L., Lebet, T., An, J.,
Sawalle-Belohradsky, J., Belohradsky, B.H., Heinz, V., Freeman, A.F., Xu, Y., Jenne, C.N., Foger, N., Sorensen, R.U., Goodnow, C.C., Bear, J.E.,
Sullivan, K.E., Holland, S.M., Torgerson, T.R., Al-Herz, W., Chou, J., Hanson, I.C., Puck, J.M., Cyster, J.G., 2008. The actin regulator coronin 1A is mutant in
Albert, M.H., Geha, R.S., Renner, E.D., Orange, J.S., 2013. Defective actin a thymic egress-decient mouse strain and in a patient with severe combined
accumulation impairs human natural killer cell function in patients with dedicator of immunodeciency. Nat. Immunol. 9, 13071315.
cytokinesis 8 deciency. J. Allergy Clin. Immunol. 131, 840848. Snapper, S.B., Rosen, F.S., Mizoguchi, E., Cohen, P., Khan, W., Liu, C.H.,
Molina, I.J., Sancho, J., Terhorst, C., Rosen, F.S., Remold-Odonnell, E., 1993. T cells Hagemann, T.L., Kwan, S.P., Ferrini, R., Davidson, L., Bhan, A.K., Alt, F.W., 1998.
of patients with the Wiskott-Aldrich syndrome have a restricted defect in prolif- Wiskott-Aldrich syndrome protein-decient mice reveal a role for WASP in T but not
erative responses. J. Immunol. 151, 43834390. B cell activation. Immunity 9, 8191.
Moshous, D., Martin, E., Carpentier, W., Lim, A., Callebaut, I., Canioni, D., Hauck, F., Snapper, S.B., Meelu, P., Nguyen, D., Stockton, B.M., Bozza, P., Alt, F.W.,
Majewski, J., Schwartzentruber, J., Nitschke, P., Sirvent, N., Frange, P., Picard, C., Rosen, F.S., Von Andrian, U.H., Klein, C., 2005. WASP deciency leads to global
Blanche, S., Revy, P., Fischer, A., Latour, S., Jabado, N., De Villartay, J.P., 2013. defects of directed leukocyte migration in vitro and in vivo. J. Leukoc. Biol. 77,
Whole-exome sequencing identies Coronin-1A deciency in 3 siblings with 993998.
immunodeciency and EBV-associated B-cell lymphoproliferation. J. Allergy Clin. Su, H.C., 2010. Dedicator of cytokinesis 8 (DOCK8) deciency. Curr. Opin. Allergy Clin.
Immunol. 131, 15941603. Immunol. 10, 515520.
Moulding, D.A., Blundell, M.P., Spiller, D.G., White, M.R., Cory, G.O., Calle, Y., Suetsugu, S., Miki, H., Takenawa, T., 1999. Identication of two human WAVE/SCAR
Kempski, H., Sinclair, J., Ancliff, P.J., Kinnon, C., Jones, G.E., Thrasher, A.J., homologues as general actin regulatory molecules which associate with the Arp2/3
2007. Unregulated actin polymerization by WASp causes defects of mitosis and complex. Biochem. Biophys. Res. Commun. 260, 296302.
cytokinesis in X-linked neutropenia. J. Exp. Med. 204, 22132224. Thrasher, A.J., Burns, S.O., 2010. WASP: a key immunological multitasker. Nat. Rev.
Moulding, D.A., Moeendarbary, E., Valon, L., Record, J., Charras, G.T., Thrasher, A.J., Immunol. 10, 182192.
2012. Excess F-actin mechanically impedes mitosis leading to cytokinesis failure in Tsuboi, S., Meerloo, J., 2007. Wiskott-Aldrich syndrome protein is a key regulator of
X-linked neutropenia by exceeding Aurora B kinase error correction capacity. Blood the phagocytic cup formation in macrophages. J. Biol. Chem. 282, 3419434203.
120, 38033811. Tsuboi, S., 2007. Requirement for a complex of Wiskott-Aldrich syndrome protein
Moulding, D.A., Record, J., Malinova, D., Thrasher, A.J., 2013. Actin cytoskeletal (WASP) with WASP interacting protein in podosome formation in macrophages. J.
defects in immunodeciency. Immunol. Rev. 256, 282299. Immunol. 178, 29872995.
Mueller, P., Massner, J., Jayachandran, R., Combaluzier, B., Albrecht, I., Gateld, J., van de Vijver, E., Maddalena, A., Sanal, O., Holland, S.M., Uzel, G., Madkaikar, M.,
Blum, C., Ceredig, R., Rodewald, H.R., Rolink, A.G., Pieters, J., 2008. Regulation of T De Boer, M., Van Leeuwen, K., Koker, M.Y., Parvaneh, N., Fischer, A., Law, S.K.,
cell survival through coronin-1-mediated generation of inositol-1,4,5-trisphosphate Klein, N., Tezcan, F.I., Unal, E., Patiroglu, T., Belohradsky, B.H., Schwartz, K.,
and calcium mobilization after T cell receptor triggering. Nat. Immunol. 9, 424431. Somech, R., Kuijpers, T.W., Roos, D., 2012. Hematologically important mutations:
Mueller, P., Liu, X., Pieters, J., 2011. Migration and homeostasis of nave T cells leukocyte adhesion deciency (rst update). Blood Cells Mol. Dis. 48, 5361.
depends on coronin 1-mediated prosurvival signals and not on coronin van de Vijver, E., van den Berg, T.K., Kuijpers, T.W., 2013. Leukocyte adhesion
1-dependent lamentous actin modulation. J. Immunol. 186, 40394050. deciencies. Hematol. Oncol. Clin. North Am. 27, 101116. viii.
Mullins, R.D., Heuser, J.A., Pollard, T.D., 1998. The interaction of Arp2/3 complex Walmsley, M.J., Ooi, S.K., Reynolds, L.F., Smith, S.H., Ruf, S., Mathiot, A., Vanes, L.,
with actin: nucleation, high afnity pointed end capping, and formation of Williams, D.A., Cancro, M.P., Tybulewicz, V.L., 2003. Critical roles for Rac1 and
branching networks of laments. Proc. Natl. Acad. Sci. U.S.A. 95, 61816186. Rac2 GTPases in B cell development and signaling. Science 302, 459462.
Mulloy, J.C., Cancelas, J.A., Filippi, M.D., Kalfa, T.A., Guo, F., Zheng, Y., 2010. Rho Welch, M.D., Iwamatsu, A., Mitchison, T.J., 1997. Actin polymerization is induced by
GTPases in hematopoiesis and hemopathies. Blood 115, 936947. Arp2/3 protein complex at the surface of Listeria monocytogenes. Nature 385,
de Noronha, S., Hardy, S., Sinclair, J., Blundell, M.P., Strid, J., Schulz, O., Zwirner, J., 265269.
Jones, G.E., Katz, D.R., Kinnon, C., Thrasher, A.J., 2005. Impaired dendritic-cell Westerberg, L., Greicius, G., Snapper, S.B., Aspenstrom, P., Severinson, E., 2001.
homing in vivo in the absence of Wiskott-Aldrich syndrome protein. Blood 105, Cdc42, Rac1, and the Wiskott-Aldrich syndrome protein are involved in the
15901597. cytoskeletal regulation of B lymphocytes. Blood 98, 10861094.
Ochs, H.D., Thrasher, A.J., 2006. The Wiskott-Aldrich syndrome. J. Allergy Clin. Westerberg, L., Larsson, M., Hardy, S.J., Fernandez, C., Thrasher, A.J.,
Immunol. 117, 725738. Severinson, E., 2005. Wiskott-Aldrich syndrome protein deciency leads to
Ochs, H.D., Slichter, S.J., Harker, L.A., Von Behrens, W.E., Clark, R.A., reduced B-cell adhesion, migration, and homing, and a delayed humoral immune
Wedgwood, R.J., 1980. The Wiskott-Aldrich syndrome: studies of lymphocytes, response. Blood 105, 11441152.
granulocytes, and platelets. Blood 55, 243252. Wetzler, M., Talpaz, M., Kleinerman, E.S., King, A., Huh, Y.O., Gutterman, J.U.,
Ochs, H.D., Filipovich, A.H., Veys, P., Cowan, M.J., Kapoor, N., 2009. Wiskott-Aldrich Kurzrock, R., 1990. A new familial immunodeciency disorder characterized by
syndrome: diagnosis, clinical and laboratory manifestations, and treatment. Biol. severe neutropenia, a defective marrow release mechanism, and hypogamma-
Blood Marrow Transplant. 15, 8490. globulinemia. Am. J. Med. 89, 663672.
Immune Deficiency j WiskottAldrich Syndrome, Leukocyte Adhesion Deficiency, and Other Migration Defects 425

Williams, D.A., Tao, W., Yang, F., Kim, C., Gu, Y., Manseld, P., Levine, J.E., Zhang, H., Schaff, U.Y., Green, C.E., Chen, H., Sarantos, M.R., Hu, Y., Wara, D.,
Petryniak, B., Derrow, C.W., Harris, C., Jia, B., Zheng, Y., Ambruso, D.R., Simon, S.I., Lowell, C.A., 2006. Impaired integrin-dependent function in Wiskott-
Lowe, J.B., Atkinson, S.J., Dinauer, M.C., Boxer, L., 2000. Dominant negative Aldrich syndrome protein-decient murine and human neutrophils. Immunity 25,
mutation of the hematopoietic-specic Rho GTPase, Rac2, is associated with 285295.
a human phagocyte immunodeciency. Blood 96, 16461654. Zhang, Q., Davis, J.C., Lamborn, I.T., Freeman, A.F., Jing, H., Favreau, A.J.,
Wiskott, A., 1937. Familiarer, angeborener Morbus Werlhoi? Monatsschr. Kinder- Matthews, H.F., Davis, J., Turner, M.L., Uzel, G., Holland, S.M., Su, H.C., 2009.
heilkd. 68, 212216. Combined immunodeciency associated with DOCK8 mutations. N. Engl. J. Med.
Worth, A., 2013. Unpublished data, personal communication. 361, 20462055.
Yamauchi, A., Kim, C., Li, S., Marchal, C.C., Towe, J., Atkinson, S.J., Dinauer, M.C., Zicha, D., Allen, W.E., Brickell, P.M., Kinnon, C., Dunn, G.A., Jones, G.E.,
2004. Rac2-decient murine macrophages have selective defects in superoxide Thrasher, A.J., 1998. Chemotaxis of macrophages is abolished in the Wiskott-
production and phagocytosis of opsonized particles. J. Immunol. 173, 59715979. Aldrich syndrome. Br. J. Haematol. 101, 659665.
Zhang, J., Shehabeldin, A., Da Cruz, L.A., Butler, J., Somani, A.K., Mcgavin, M., Zuchero, J.B., Coutts, A.S., Quinlan, M.E., Thangue, N.B., Mullins, R.D., 2009.
Kozieradzki, I., Dos Santos, A.O., Nagy, A., Grinstein, S., Penninger, J.M., p53-cofactor JMY is a multifunctional actin nucleation factor. Nat. Cell Biol. 11,
Siminovitch, K.A., 1999. Antigen receptor-induced activation and cytoskeletal 451459.
rearrangement are impaired in Wiskott-Aldrich syndrome protein-decient
lymphocytes. J. Exp. Med. 190, 13291342.

Вам также может понравиться