peeking into the secret life of neutrophils

14
SINGAPORE IMMUNOLOGY NETWORK Peeking into the secret life of neutrophils Jackson LiangYao Li Lai Guan Ng Published online: 11 March 2012 Ó Springer Science+Business Media, LLC 2012 Abstract The migration of neutrophils between tissue compartments is an important aspect of innate immune surveil- lance. This process is regulated by a cascade of cellular and molecular signals to avoid unnecessary crowding of neu- trophils at the periphery, to allow rapid mobilization of neutrophils in response to inflammatory stimuli, and to return to a state of homeostasis after the response. Intravital microscopy approaches have been fundamental in unraveling many aspects of neutrophil behavior, providing important mechanistic information on the processes involved in basal and disease states. Here, we provide a broad overview of the current state of research on neutrophil biology, describing the processes in the typical life cycle of neutrophils, from their first appearance in the bone marrow until their eventual destruction. We will focus on novel aspects of neutrophil behavior, which had previously been elusive until their recent elucidation by advanced intravital microscopy techniques. Keywords Neutrophil Á Intravital microscopy Á Bone marrow Á Blood Á Vessel Á Interstitium Á Effector Á Clearance Á Basal Á Inflammation Introduction Neutrophils represent a major subset of innate immune cells and play a key role in the defense against invading pathogens. It is well established that neutrophils are rapidly recruited into sites of inflammation or infection, where they can destroy bacterial and fungal pathogens by several mechanisms including the release of enzymes and reactive oxygen species, as well as through direct phagocytosis. Proteolytic enzymes and oxidants released by neutrophils can however also cause collateral damage to the sur- rounding tissue. Due to the highly destructive nature of these cells, neutrophil homeostasis is maintained through tight regulation of neutrophil production in the bone marrow (granulopoiesis), distribution between different compartments, and migration through interstitial tissues. One of the most fundamental characteristics of neutro- phils is their ability to shuttle between different body compartments in response to inflammatory stimuli. Under homeostatic conditions, bone marrow serves as the major reservoir for neutrophils, and only 1–2% of neutrophils are present in the circulation in mice [1]. During acute inflammatory responses, neutrophils can be rapidly released from bone marrow into the bloodstream, resulting in a dramatic increase in circulating neutrophil numbers. To enter sites of inflammation in the tissue, neutrophils need to move across the blood vessel walls. The cascade of events leading to neutrophil extravasation has been exam- ined extensively, and these studies have provided a wealth of information about the mechanisms underlying the recruitment of neutrophils to sites of inflammation. In contrast, the subsequent cellular and molecular events involved in neutrophil migration within interstitial tissue are not well defined. This aspect of neutrophil mobilization has been overlooked due, at least partly, to the lack of J. L. Li Á L. G. Ng (&) Singapore Immunology Network (SIgN), Agency for Science, Technology and Research (A*STAR), 8A Biomedical Grove, #03 Immunos, Biopolis, Singapore 138648, Singapore e-mail: [email protected] Lai Guan Ng 123 Immunol Res (2012) 53:168–181 DOI 10.1007/s12026-012-8292-8

Upload: lai-guan-ng

Post on 25-Aug-2016

213 views

Category:

Documents


1 download

TRANSCRIPT

  • SINGAPORE IMMUNOLOGY NETWORK

    Peeking into the secret life of neutrophils

    Jackson LiangYao Li Lai Guan Ng

    Published online: 11 March 2012

    Springer Science+Business Media, LLC 2012

    Abstract The migration of neutrophils between tissue compartments is an important aspect of innate immune surveil-

    lance. This process is regulated by a cascade of cellular and molecular signals to avoid unnecessary crowding of neu-

    trophils at the periphery, to allow rapid mobilization of neutrophils in response to inflammatory stimuli, and to return to a

    state of homeostasis after the response. Intravital microscopy approaches have been fundamental in unraveling many

    aspects of neutrophil behavior, providing important mechanistic information on the processes involved in basal and disease

    states. Here, we provide a broad overview of the current state of research on neutrophil biology, describing the processes in

    the typical life cycle of neutrophils, from their first appearance in the bone marrow until their eventual destruction. We will

    focus on novel aspects of neutrophil behavior, which had previously been elusive until their recent elucidation by advanced

    intravital microscopy techniques.

    Keywords Neutrophil Intravital microscopy Bone marrow Blood Vessel Interstitium Effector Clearance Basal Inflammation

    Introduction

    Neutrophils represent a major subset of innate immune

    cells and play a key role in the defense against invading

    pathogens. It is well established that neutrophils are rapidly

    recruited into sites of inflammation or infection, where they

    can destroy bacterial and fungal pathogens by several

    mechanisms including the release of enzymes and reactive

    oxygen species, as well as through direct phagocytosis.

    Proteolytic enzymes and oxidants released by neutrophils

    can however also cause collateral damage to the sur-

    rounding tissue. Due to the highly destructive nature of

    these cells, neutrophil homeostasis is maintained through

    tight regulation of neutrophil production in the bone

    marrow (granulopoiesis), distribution between different

    compartments, and migration through interstitial tissues.

    One of the most fundamental characteristics of neutro-

    phils is their ability to shuttle between different body

    compartments in response to inflammatory stimuli. Under

    homeostatic conditions, bone marrow serves as the major

    reservoir for neutrophils, and only 12% of neutrophils are

    present in the circulation in mice [1]. During acute

    inflammatory responses, neutrophils can be rapidly

    released from bone marrow into the bloodstream, resulting

    in a dramatic increase in circulating neutrophil numbers.

    To enter sites of inflammation in the tissue, neutrophils

    need to move across the blood vessel walls. The cascade of

    events leading to neutrophil extravasation has been exam-

    ined extensively, and these studies have provided a wealth

    of information about the mechanisms underlying the

    recruitment of neutrophils to sites of inflammation. In

    contrast, the subsequent cellular and molecular events

    involved in neutrophil migration within interstitial tissue

    are not well defined. This aspect of neutrophil mobilization

    has been overlooked due, at least partly, to the lack of

    J. L. Li L. G. Ng (&)Singapore Immunology Network (SIgN), Agency for Science,

    Technology and Research (A*STAR), 8A Biomedical Grove,

    #03 Immunos, Biopolis, Singapore 138648, Singapore

    e-mail: [email protected]

    Lai Guan Ng

    123

    Immunol Res (2012) 53:168181

    DOI 10.1007/s12026-012-8292-8

  • suitable technologies to address this type of questions in

    the past. It is also possible that neutrophils were typically

    considered to be too short-lived to play a significant role in

    the immune response after entering interstitial tissues (few

    hours) [2, 3]. However, a recent study, which used deute-

    rium to label neutrophils in vivo, has shown that non-

    activated neutrophils may have a longer lifespan than

    previously thought ([half a day in mice, 5.4 days inhumans) [4]. In contrast to the general belief that neutro-

    phils can only migrate into tissues in response to inflam-

    matory stimuli, we recently showed that neutrophils are

    present in the interstitial tissue of the skin even at resting

    state [5]. This indicates that neutrophils may have an

    important role in maintaining tissue homeostasis and

    shaping immune responses. Thus, a better understanding of

    neutrophil behavior and function in living tissues may

    reveal new knowledge about neutrophil biology.

    Historically, intravital microscopy (IVM) techniques

    have been instrumental in unraveling the complex mech-

    anisms of neutrophil behavior. The first description of the

    leukocyte extravasation process was provided by an IVM

    study performed more than 100 years ago [6]. The con-

    tinued advancement in microscopy technology coupled

    with the availability of a wide array of genetically modified

    mice has helped to further define the molecular basis of

    neutrophil migration. One microscopy approach that has

    received a lot of attention in recent years is multiphoton

    (MP) microscopy. This technique has overcome previous

    limitations of conventional microscopy by allowing a

    greater depth of penetration as a result of its localized

    nonlinear laser signal generation. As such, it is now rou-

    tinely used to perform dynamic, multidimensional imaging

    to simultaneously track cell populations at the single-cell

    level in living tissues or organs. For more detailed tech-

    nical information about the general setup of an MP

    microscope, please refer to these comprehensive review

    articles [7, 8].

    The availability of new technology will often open up

    new avenues of research, and this is certainly true in the

    case of MP microscopy. Immunologists have long adopted

    IVM techniques to address biological questions relating to

    leukocyte trafficking. However, IVM approaches using

    linear-absorption microscopy techniques such as bright-

    field, epifluorescence, and laser scanning confocal

    microscopy are limited to the imaging of translucent tissues

    or superficial regions of non-translucent organs due to the

    lack of tissue penetration. Since the first application of MP

    imaging for the study of immune cells in intact organs or

    tissues in early 2000 [912], this technology has become an

    increasingly important tool in the field of immunology.

    Notably, adaptive immune cells are relatively better-char-

    acterized by MP imaging studies than innate immune cells.

    T cells arguably represent the most studied immune cells

    by MP imaging, and important information has been

    obtained about their behavior at resting stage and during

    various pathological settings within their native microen-

    vironment (e.g., in tumors and infections, see reviews

    [1315]). In contrast, MP microscopy has only recently

    been used to study neutrophil dynamics in vivo. Despite

    this, these studies have already provided us with important

    clues as to how neutrophil migration is regulated. In this

    review, we will provide a broad description of neutrophil

    behavior and their immune surveillance strategies, includ-

    ing granulopoiesis, trafficking, adhesion, crawling, trans-

    migration, swarming, effector functions, clearance, among

    others. In particular, we will focus on discussing the results

    obtained from IVM imaging of neutrophils (especially

    those from MP microscopy) in various mouse models,

    highlighting the new perspectives offered by these studies.

    The behavior of neutrophils in different tissue

    compartments

    Bone marrow

    Neutrophils make their first appearance in the hematopoi-

    etic cords of the bone marrow [16]. A series of proliferating

    myeloid precursors generate non-dividing neutrophilic

    metamyelocytes, which then differentiate linearly into

    mature segmented neutrophils [17, 18] (See Fig. 1). Pro-

    duction and maturation of neutrophils appear to be con-

    fined to the bone marrow, and this is in line with the role of

    neutrophils as the major innate effector cells of the immune

    systemunlike B or T lymphocytes, which require highly

    specialized training in additional tissues for maturation.

    Within the bone marrow, neutrophils are the most abundant

    leukocytes producedapproximately 107 neutrophils are

    generated each day in mice [19], and 1011 neutrophils per

    day in humans [2]. A high constitutive neutrophil produc-

    tion rate is important because an effective neutrophil

    response, when initiated, typically requires the concerted

    action of very large numbers of neutrophilsa high total

    threshold number thus needs to be maintained even during

    physiological basal conditions. Production of neutrophils

    takes approximately 2.3 days to complete in mice, with at

    least 16 h spent in the mitotic pool (promyelocytes and

    myelocytes) [4], and thus, requirements for neutrophil

    supply during an infectious episode are unlikely to be

    achieved without a large reserve pool. In addition, neu-

    trophils have a relatively short half-life compared to other

    leukocytes, thus necessitating high production rates to

    maintain a constant turnover [19]. The short-lived nature of

    neutrophils might be an evolutionary defense against the

    development of parasites (that might otherwise adapt to

    colonizing neutrophils) [2022], but may simply reflect a

    Singapore Immunology Network: SIgN (2012) 53:168181 169

    123

  • more stringent functional requirement for neutrophil

    maintenance [23] (e.g., since the unplanned necrosis of

    aging neutrophils in the periphery would release noxious

    mediators that are highly damaging to the surrounding

    tissue).

    The bone marrow acts as an important reservoir to hold

    excess neutrophils, maintaining the high total threshold

    number of neutrophils and preventing the unnecessary

    crowding of neutrophils in the periphery. At the physio-

    logical state in mice, it is estimated that more than 90% of

    the total neutrophils are held within the bone marrow, with

    only 12% present in the circulation [1, 24]. The majority

    of mature neutrophils localize in the extravascular com-

    partments between the sinusoids of the bone marrow. A

    small proportion of neutrophils (30%) are motile and crawl

    randomly within the space (mean migration veloc-

    ity = 1.5 lm/min) [25]. An even smaller percentage ofneutrophils may spontaneously exit the bone marrow by

    entering the surrounding sinusoids, accounting for the basal

    blood neutrophil counts. However, upon acute tissue injury

    or other forms of inflammation, endothelia or other

    immune cells at the peripheral site such as macrophages

    may release granulocyte colony-stimulating factor (G-

    CSF), a glycoprotein growth factor that maintains granu-

    locyte survival and increases granulocyte proliferation

    (granulopoiesis) [16]. In addition to stimulating granulo-

    poiesis, G-CSF also functions as a mobilizing agent that

    triggers the release of bone marrow neutrophils into the

    bloodstream [1]. Despite the well-established role of

    G-CSF in neutrophil mobilization from the bone marrow,

    the spatiotemporal regulation of this process remains

    poorly defined. Recently, an MP-IVM study by Gunzers

    laboratory [25] showed that a single dose of G-CSF

    intraperitoneally injected in mice triggered neutrophil

    mobilization by increasing both neutrophil motility and

    directionality in the bone marrow. Mobilized neutrophils

    were found to migrate toward the nearest sinusoids and exit

    the bone marrow, leading to a dramatic increase in the

    number of circulating neutrophils. Blood neutrophil counts

    were found to peak 2 h after G-CSF injection, while the

    increased neutrophil mean migration velocity in the bone

    marrow persisted for up to 4 h post-injection, and the

    increased proportion of motile neutrophils persisted for

    greater than 16 h.

    The mechanisms regulating the number of neutrophils

    being released from the bone marrow has been proposed to

    involve the chemokine receptors CXCR4 and CXCR2

    (IL8-R beta) expressed on neutrophils, in which CXCR4 is

    important for the retention of neutrophils in the bone

    marrow, whereas CXCR2 is important for neutrophil

    release from the bone marrow [26]. The major binding

    partner for CXCR4 is CXCL12 (SDF-1) [27], which is

    expressed by reticular cells [28] of the bone marrow. On

    the other hand, the major binding partners for CXCR2 are

    CXCL1 (KC) and CXCL2 (MIP-2a), which can be secretedby endothelial cells [29]. Neutrophil retention in the bone

    marrow is believed to be dependent on CXCR4CXCL12

    interactions between neutrophils and bone marrow reticular

    cells. Consistent with this hypothesis, neutrophil numbers

    are reduced in the bone marrow and elevated in the

    peripheral blood of CXCR4-deficient mice [30], and anti-

    CXCR4 blocking antibodies have been shown to mobilize

    Fig. 1 A schematicrepresentation of neutrophil

    development during

    granulopoiesis. Top right Az-projection multiphoton imageof skull bone marrow in

    Lysozyme-GFP C57BL/6

    mouse injected with TRITC

    dextran for blood vessel

    labeling. Skull bone collagen

    shows up as second harmonic

    generation (SHG). Image

    courtesy of Dr. Yilin Wang

    (SIgN, Singapore). Left andbottom Development ofneutrophils in the bone marrow,

    showing maturation stages

    commonly described in

    hematology. Relationships with

    other hematopoietic cells are

    also shown

    170 Singapore Immunology Network: SIgN (2012) 53:168181

    123

  • bone marrow neutrophils [31]. Previously, G-CSF was

    proposed to mobilize neutrophils from the bone marrow by

    activating proteases that cleave CXCL12 [32]. However, a

    study using protease-deficient mice demonstrated that these

    mice retain the ability to mobilize bone marrow neutrophils

    [33], and another study showed the down-regulation of

    both CXCR4 and CXCL12 by G-CSF [34], providing a

    simpler explanation of the relationship between G-CSF and

    CXCR4-mediated retention. On the other hand, CXCR2

    activation by its ligands CXCL1 or CXCL2 is proposed to

    be an important chemotactic signal for neutrophil emigra-

    tion. G-CSF injections were unable to mobilize neutrophils

    from the bone marrow of CXCR2-deficient mice, thus

    demonstrating that G-CSF-induced mobilization is cen-

    trally dependent on CXCR2 [25]. In agreement with this

    observation, CXCR2 ligands are elevated in blood during

    G-CSF-induced mobilization, which thus provides a simple

    mechanistic explanation for neutrophil mobilization from

    bone marrow into the sinusoids. During sepsis, however,

    CXCL12 concentrations become elevated in blood [35],

    providing a reverse gradient described earlier for CXCR4

    CXCL12 binding retention, indicating that the actual

    mechanisms of regulation may be more complex than

    conceptualized here.

    Perhaps of more novelty is the discovery that megakary-

    ocytes, the precursors of platelets, are able to produce

    CXCR2 ligands in response to G-CSF, mediated by throm-

    bopoietin (TPO) [25]. Megakaryocytes, which derive from a

    common myeloid progenitor, reside in the bone marrow and

    are usually found lining the sinusoids. An MP imaging study

    from von Andrians laboratory [36] has provided the first

    dynamic view of megakaryocyte activities in vivo. The

    authors showed that perisinusoidal megakaryocytes were

    able to send out cytoplasmic projections through the vessel

    walls that break off from their transendothelial stems to form

    platelets. Although not conclusively proven, the fact that

    portions of the megakaryocytes cross the vessel walls sug-

    gests the enticing idea that these cells can release CXCL1

    and/or CXCL2 directly into the bloodstream in very close

    vicinity to bone marrow neutrophils and thus provide or

    regulate directional cues for CXCR2-mediated mobilization

    [25] (See Fig. 2). This might help to explain the mechanism

    behind the seemingly stochastic neutrophil mobilization

    during basal state, since systemic CXCR2 ligands generated

    from a distant source would be expected to quickly dilute in a

    homogeneous manner such that chemokine gradients would

    be detected almost equally well by the majority of bone

    marrow neutrophils.

    Blood

    Neutrophils patrol the body by circulating in the blood,

    which allows them to respond quickly to signals of

    pathogen entry or loss of tissue integrity. To detect

    pathogens, neutrophils express innate pathogen recognition

    receptors (PRRs), of which Toll-like receptors (TLRs)

    feature prominently [37]. These receptors detect evolu-

    tionarily conserved signature molecules of commonly

    encountered pathogens, which are collectively known as

    pathogen-associated molecular patterns (PAMPs). Neutro-

    phils express many of the known TLRs [38] and, upon

    activation, typically undergo pro-inflammatory changes,

    including increased propensities for phagocytosis and

    cytokine secretion [39]. For detecting tissue injury, analo-

    gous damage-associated molecular patterns (DAMPs) [37]

    are thought to mediate inflammatory responses of neutro-

    phils. DAMPs are molecules associated with necrotizing

    cells, and examples include ATP, uric acid, heat-shock

    proteins, and mitochondrial DNA [40]. At basal level, there

    are approximately 5001,000 neutrophils per microliter of

    blood, although the actual numbers can vary among mouse

    strains [41]. This basal level of neutrophils in the periphery

    allows prompt responses since the earliest neutrophils can

    arrive at the target site without the need to travel vast

    distances from the bone marrow in order to establish the

    inflammatory response. This also allows a graded response

    to injury or pathogen, since minor insults need only trigger

    a localized response that draws nearby circulating neutro-

    phils, without unnecessarily mobilizing neutrophils from

    the bone marrow. The surveillance program of neutrophils

    takes advantage of the ubiquitous perfusion of blood

    throughout the body, allowing neutrophils to extend their

    reach of response into most tissues, although immune-

    privileged sites such as the brain have other specialized cell

    types (i.e., microglia and astrocytes) to carry out similar

    Fig. 2 Relationship between CXCR2 and CXCR4 in neutrophilmobilization and retention in bone marrow. Neutrophils are retained

    in the bone marrow by the interaction between CXCR4 and its only

    known ligand CXCL12. During mobilization, GCSF reduces the

    strength of this interaction by down-regulating CXCR2 and CXCL12

    expression levels, while CXCR2 ligands in the blood, which may

    include CXCL1 and CXCL2, provide the chemokine gradients for

    neutrophil extravasation into sinusoids. Also shown is the hypothet-

    ical contribution of CXCR2 ligands by megakaryocytes into the

    bloodstream, as megakaryocytes have been shown to be able to

    express them under GCSF stimulation via TPO

    Singapore Immunology Network: SIgN (2012) 53:168181 171

    123

  • duties [42]. In accordance with this theory, an MP-IVM

    study in mice from Dustins laboratory showed that laser

    injury to the brain parenchyma triggered the rapid exten-

    sion of microglia dendrites and polarization of the astrocyte

    cytoplasm toward the injury site, without causing any

    neutrophil infiltration [43].

    Peripheral neutrophils are also known to spend addi-

    tional time transiting certain highly vascularized organs

    such as the lungs and liver [44], although it is unclear

    whether this reflects an increased need for the patrolling of

    these regions or whether these organs represent secondary

    storage depots for additional fine-tuning of blood neutro-

    phil counts. In fact, large numbers of extravascular neu-

    trophils can be observed in healthy lung tissue [45] despite

    the fact that neutrophils are thought to rarely exit blood

    vessels except during activation. Using a combination of

    MP-IVM and functional approaches, we have recently

    shown that even under homeostatic conditions, a very small

    percentage of neutrophils can be found in the mouse skin

    interstitium, as well as within lymph fluid (demonstrated in

    sheep) [5]. This perhaps signifies that the neutrophil patrol

    routes incorporate not only the blood circulation but

    interstitial and lymphoid paths as well.

    However, it is clear that neutrophils primarily employ

    blood vessels of healthy tissue as highways to access various

    locations of injury or pathogen detection in the interstitium. In

    mice, blood is pumped at approximately 10 beats per second

    [46] and can flow at approximately 1 mm/s in a mid-sized

    venule [47]this velocity represents approximately a hun-

    dred neutrophil lengths each second. While such high speeds

    are useful for transporting neutrophils rapidly from site to site,

    this compromises the ability of neutrophils to sense and

    respond to subtle cues presented on the vascular walls. Thus,

    upon activation, neutrophils must exit the central bloodstream

    and roll along the vessel walls in the boundary layer (where

    shear forces dictate lower fluid flow velocities) in order to

    sample the more precise chemotactic signals. Under basal

    conditions in the skin, blood vessels are shown to express

    CD62E and CD62P (E- and P-selectins) [48], which allow

    neutrophils to exhibit constitutive rolling. This may represent

    a more nuanced patrolling strategy for skin, given the high

    correlation between the loss of its integrity and subsequent

    pathogen invasion, although it is unclear whether other sites at

    high risk of pathogen influx also exhibit constitutive neutro-

    phil rolling [49]. In a large venule (*30 lm diameter), con-stitutive rolling in the skin occurs at approximately 10 lm persecond, with approximately 50% of total leukocytes display-

    ing such behavior [50].

    Blood vessel walls

    Neutrophils are able to respond to pathogens and tissue

    injury through specialized receptors described earlier, but

    as the majority of patrolling neutrophils are found flowing

    in the blood, they are usually unable to directly make

    contact with the pathogen or injured tissue. As such, they

    often depend on other sentinel cells to provide the distress

    signals for their recruitment. Many cell types can trigger

    neutrophil recruitment directly or indirectly. In the skin,

    these cells can be endothelial cells or other sentinel cells

    such as dendritic cells [51, 52], mast cells [53], macro-

    phages [54], basophils [55], and monocytes [56], with the

    actual mechanisms used varying according to the context.

    Endothelial cells of the blood vessel walls constitute a very

    important sentinel cell type, but are generally overlooked

    due to their non-leukocyte nature, and are often assumed to

    be passive scaffolding for neutrophil adhesion. Impor-

    tantly, however, they have been shown to express classical

    molecules involved in pathogen detection (which are more

    often associated with leukocytes), including CD14, TLR2,

    TLR4, TLR9, MD2, and MyD88 [49], and thus provide a

    large surface area for pathogen detection, which can lead to

    neutrophil activation. Endothelial cells are also crucially

    important for neutrophil function, since neutrophils require

    their active cooperation for successful tethering, rolling,

    adhesion, crawling, and transmigration in order to exit the

    bloodstream and travel to the target site.

    Since chemokine signals that enter the blood would be

    quickly flushed away and potentially scavenged (e.g., by

    DARCs on erythrocytes), it was somewhat of a mystery as

    to how a neutrophil could recognize when and where to

    exit the blood stream in order to migrate to the injury site.

    A seminal study by Kubess laboratory [57] used a sterile

    hepatic focal injury IVM model in which well-defined

    thermal injuries could be generated and examined the

    subsequent neutrophil recruitment processes from liver

    sinusoids. They found that contrary to common perception,

    ATP (in its role as a DAMP) did not function in vivo to

    trigger direct neutrophil activation, and neither did it serve

    as a chemoattractant. Instead, ATP triggered inflamma-

    some-dependent cell signaling processes within the

    microenvironment that culminated in the activation of

    endothelial cells, which upregulated CD54 (ICAM-1) to

    provide neutrophils with a surface conducive for activation

    and recruitment. Modifications of glycocalyx thickness

    might also be necessary to expose the adhesion molecules

    on the endothelia cell surface [58]. In addition, they also

    found that CXCL2 formed a gradient on the luminal side of

    the blood vessel walls and was mainly responsible for

    CXCR2-mediated neutrophil chemotaxis there. CXCL1

    played a minor role in this process. The chemokine gra-

    dient was made possible by the immobilization of che-

    mokines on the vessel walls, presumably mediated by

    endothelial surface heparan sulfate [59]. However, this

    gradient started abruptly approximately 150 lm from theinjury site although proximal endothelial cells were still

    172 Singapore Immunology Network: SIgN (2012) 53:168181

    123

  • intact. Despite the lack of a CXCL2 gradient, neutrophils

    were still able to travel into the center of the necrotic tissue.

    This was attributed to a second gradient of formylated

    peptides that overrode the CXCL2 signal [57]. Formylated

    peptides are DAMPs of mitochondrial origin recognized by

    formyl-peptide receptor 1 (FPR-1) [60]. Thus, the current

    paradigm on neutrophil recruitment consists of a series of

    sequential signals that act at different ranges, and these

    signals may progressively override earlier signals [6164]

    (e.g., G-CSF at ultra-long range, ATP at mid-range via

    endothelial cells, CXCL2 at short range, and formylated

    peptides at immediate range)neutrophils thus start off

    heading in the general direction where they are needed, but

    once they get closer, they make decisional switches and

    respond to progressively more precise cues in order to

    finally arrive at the intended site. Recently, hydrogen per-

    oxide was shown to function as a tissue-scale chemoat-

    tractant in the zebrafish for the recruitment of leukocytes to

    injury [65] and thus might represent another layer of sig-

    naling for neutrophils, but it remains to be seen whether

    hydrogen peroxide also has a function in the mammalian

    system.

    As previously mentioned, in order to cross the blood

    vessel walls from high initial flow velocity, activated

    neutrophils have to slow down and come to a stop before

    transmigration can take place. This process, commonly

    termed the leukocyte adhesion cascade (since similar pro-

    cesses also occur for other leukocytes), has been exten-

    sively studied [66], but many of the details have only been

    elucidated relatively recently, with the advent of IVM

    technology. The initial step triggering the slowing down of

    fast-flowing neutrophils is termed tethering or capture and

    is largely mediated by selectins expressed on both the

    neutrophils and the endothelial cells, which have corre-

    sponding ligands on the opposing cell for attachment.

    CD162 (PSGL-1) is an important selectin ligand expressed

    on the neutrophil that interacts with CD62E and CD62P

    (E- and P-selectin) expressed on the endothelial cells [49].

    However, in the scenario that immune complexes form and

    deposit on the endothelium, neutrophils are also able to

    employ a selectin-independent tethering process through

    the use of their constitutively expressed Fc-gamma recep-

    tor IIIB in a low-affinity binding interaction with immune

    complexes [67]. The tethering process is thus the initial

    contact between the neutrophil and the vessel wall that

    allows the neutrophil to slow down and marginalize to the

    vessel wall where rolling can then take place. During

    rolling, the decrease in velocities allows neutrophils to

    come into stronger contact with the endothelial cells and

    further decelerate. Neutrophils then activate integrins to

    mediate strong binding interactions with the integrin

    ligands presented on the endothelial cells for firm adhesion

    and crawling.

    For neutrophils, the most important integrins are CD11a/

    CD18 (LFA-1) and CD11b/CD18 (Mac-1), and both inte-

    grins bind CD54 (ICAM-1). Previously, CD11a and

    CD11b, both of which partner CD18, were thought to

    possess similar and overlapping roles in neutrophil adhe-

    sion. An elegant study by Hentzen et al. [68] over a decade

    ago using in vitro flow chamber assays and mathematical

    modeling proposed that CD11a and CD11b function

    sequentially for neutrophil initial capture and subsequent

    adhesion, based on their capture efficiency and adhesion

    stability kineticswhile CD11a was much more efficient

    than CD11b in binding CD54, CD11b was able to maintain

    that binding for a much longer period of time. It is

    important to note that the calculations of capture efficiency

    were based on end-point measurements determined by flow

    cytometry, and the authors were thus unable to place the

    relevance of their findings in the scheme of the exact

    adhesion cascade. The roles of CD11a and CD11b in the

    adhesion cascade were later validated by Phillipson et al.

    [69] in IVM experiments on the cremaster muscle in vivo.

    However, rather than playing sequential roles during the

    initial tethering and adhesion process, CD11a and CD11b

    were found to be crucial for the steps of firm adhesion and

    subsequent intraluminal crawling, respectively. In their

    study, blocking CD11a on neutrophils resulted in a pro-

    nounced inability for these cells to adhere, but those few

    that managed to adhere displayed normal crawling on the

    vessel walls. In contrast, when CD11b was blocked, neu-

    trophils could still adhere normally but were unable to

    perform intraluminal crawling. Interestingly, in monocytes,

    CD11a/CD18 is known to additionally bind CD102

    (ICAM-2) and has been demonstrated to be important for

    intraluminal crawling [70], but curiously, neutrophils uti-

    lize only CD11b/CD18 for this purpose. Antibody block-

    ade of CD54 in wild-type mice still allowed neutrophils to

    adhere, but not crawl [69], and this suggests that CD11a/

    CD18 may also utilize CD102 as a substitute for CD54

    during firm adhesion. Other molecules may also be relevant

    in certain contexts. For example, in an in vivo Escherichia

    coli infection model, neutrophils in the liver exhibited

    dependence on CD44 for adhesion instead [71].

    Activated neutrophils almost always display intraluminal

    crawling behavior after firm adhesion. One explanation

    would be the need for the neutrophils to follow the chemo-

    kine gradients in order to exit the blood vessels at the location

    closest to the injury or pathogen site, and indeed, this was

    observed in the hepatic focal injury model earlier described

    [57]. However, in many other cases, most neutrophils

    already land at locations very close to their eventual emi-

    gration site and do not appear to follow intravascular che-

    motactic gradients. Instead, this crawling behavior appears

    to be necessary for a more mundane reasonthe sampling of

    the blood vessel wall for an optimal endothelial cellcell

    Singapore Immunology Network: SIgN (2012) 53:168181 173

    123

  • junction for emigration. In the CD11b-deficient mice

    described earlier, neutrophils were still able to transmigrate

    after adhesion, but they did so with greatly reduced effi-

    ciency [69], presumably since they were forced to exit in situ

    at non-optimal sites without the ability to crawl. Indeed,

    these neutrophils were later found to transmigrate predom-

    inantly through transcellular [72] routes (directly through the

    center of an endothelial cell) [73], which would be expected

    to be less efficient than paracellular routes (via the junctions

    at the borders of endothelial cells). In line with this thinking,

    a recent dynamic imaging study showed that the main route

    of neutrophil transmigration during inflammation was

    paracellularonly approximately 10% of neutrophils

    transmigrated transcellularly in their model [74]. Interest-

    ingly, another study demonstrated that after firm adhesion,

    most neutrophils start spontaneous crawling perpendicular to

    the blood flow [75]. This perpendicular movement was

    observed only under conditions of fluid flow and could be

    reproduced in in vitro experiments without the need for

    chemokines [75], suggesting that this behavior was me-

    chanotactic in natureshear stress was necessary for the

    neutrophil to orient itself with respect to blood flow in order

    to crawl perpendicularly on the vessel wall. The perpendic-

    ular movement allows neutrophils to quickly sample the

    vessel wall for the longitudinal border junctions between two

    endothelial cells, so that efficient paracellular transmigration

    can occur in vivo. Perpendicular sampling is more efficient

    because endothelial cells under shear flow conditions form

    vessel walls in a brick-like manner, such that their width

    (perpendicular to the vessel) is only one-fifth that of their

    length (parallel to the vessel) [75]. Neutrophils are thus more

    likely to arrive at a cellcell junction traveling perpendicu-

    larly than traveling longitudinally along the length of an

    endothelial cell. In fact, the authors also found that after

    reaching their first cellcell junction through perpendicular

    movement, most neutrophils switch crawling direction to

    follow the junctions parallel to the vessel and were able to

    travel against blood flow just as easily as traveling with blood

    flow. In the same study, Vav-1 deficient neutrophils, which

    display an inability to perform perpendicular crawling or

    travel against blood flow, were also found to have impaired

    transmigration efficiency, presumably due to their inability

    to find optimal endothelial junctions for transmigration. The

    fact that neutrophils continue to crawl longitudinally after

    having already found an endothelial cellcell junction indi-

    cates that certain portions along the endothelial borders must

    provide even more optimal emigration. An early in vitro

    study found that neutrophils preferentially transmigrate at

    tricellular endothelial junctions and might thus represent the

    optimal emigration sites [76].

    Emigration, which involves the traversing of large cells

    across an essentially gap-free wall, is a highly complex

    process and involves several classes of molecules such as

    CD31 (PECAM-1), CD99, ICAMs, cadherins, junctional

    adhesion molecules, and integrins [77]. For example,

    directionality of transmigration was recently found to be

    dependent on JAM-C [74]. Despite this complexity, one

    picture that emerges from studies on transmigration is that

    the endothelial cells are not just passive gate-keepers, but

    are actively involved in getting the neutrophil across [77].

    As mentioned earlier, transmigration may occur through

    paracellular or transcellular routes. In both routes, neutro-

    phil adhesion induces the formation of endothelial adhesive

    platforms (EAPs), which are pro-adhesive domains on the

    endothelial surface [78]. Projections from the endothelial

    cell membrane, termed docking structures, form and

    facilitate subsequent transmigration [79]. During the

    paracellular route, in addition to the forces generated by the

    neutrophils when squeezing through the tight junctions,

    endothelial cells rapidly alter their cell surface adhesion

    molecule profiles to weaken endothelial cellcell attach-

    ment in order to smooth the way for neutrophil passage and

    will then rapidly re-establish these junctions once the

    neutrophils have crossed [80]. This is necessary to maintain

    the integrity of the endothelial barrier, but during severe

    inflammatory responses, junctional disruptions may ensue

    and result in increased vascular permeability, potentially

    causing edema [77]. In the case of the transcellular route,

    endothelial cells have to play an even more active role, as it

    would be inconceivable for neutrophils to punch holes

    through the endothelial cell centers without causing severe

    losses in vascular integrity. Leukocytes are proposed to

    form short actin-rich protrusions (termed podosomes) at

    their contact regions with the endothelial cell, which results

    in corresponding deformations in the endothelial cell

    membrane [81]. This has been proposed to be a probing

    mechanism for the leukocytes to identify suitable trans-

    migration spots [82]. Next, endothelial cells form migra-

    tory cups [83] that essentially allow neutrophils to sink

    into the endothelial cell. This process is analogous to a

    sequential phagocytosis and exocytosis process, because

    endothelial dome structures [73] immediately close behind

    the neutrophils after their initial entry into the endothelial

    cell, preserving the integrity of the endothelial wall before

    membrane closure after neutrophil exit (see Fig. 3).

    Interstitium

    Perhaps one of the biggest recent advances in the field of

    neutrophil research is the ability to directly visualize neu-

    trophil activities within the interstitium by MP-IVM. Using

    this approach, several studies have revealed that a charac-

    teristic behavior of neutrophils is their propensity to form

    dynamic swarms. This behavior has been characterized in

    pathogen infection models, including Toxoplasma gondii in

    the lymph node [84] and Listeria monocytogenes in the lung

    174 Singapore Immunology Network: SIgN (2012) 53:168181

    123

  • [45], as well as in sterile models, including laser injury model

    in ear dermis [5] and ischemiareperfusion injury in lung

    explants [45], indicating that swarming may be a general

    neutrophil effector strategy. Typically, multiple neutrophils

    will enter the interstitium in response to a specific stimulus

    and swarm around the stimulus, for example parasite or

    injury location. Of note, in the ischemic-reperfusion injury

    model, tissue damage is expected to be uniform, yet

    swarming behavior was still observed [45]. In the T. gondii

    model, swarms may be either small and transient or large and

    persistent [84]. Small swarms were observed to contain

    approximately 150 neutrophils and last for 1040 min before

    dissipating. This dissipation occurred as a result of the neu-

    trophils leaving the swarm to join other swarm clusters. On

    the other hand, when a swarm consisted of more than 300

    neutrophils, they grew larger in size instead of dissipating

    and remained stable for extended periods of time up to sev-

    eral hours. This increase in size could be due to the continued

    migration of neutrophils into the swarm cluster, or it could

    also arise from their merging with smaller nearby swarms.

    Large clusters may include up to 2,000 neutrophils. This

    behavior suggests a mechanism whereby neutrophils secrete

    their own chemotactic signals to induce swarming in a

    positive feedback cycle, where a certain threshold concen-

    tration exists that allows swarm numbers to be stably

    maintained [84]. In the L. monocytogenes model, the authors

    detected small, transient swarms that dissipated in

    1020 min [45], but also observed the dynamic nature of the

    swarms in which clusters grow and shrink.

    In the sterile injury model of mouse ear skin, neutrophils

    formed a single cluster at the localized laser injury site, and

    the cluster remained stable for at least an hour [5]. In this

    model, a three-phase cascade of cluster formation was

    observed, which included scouting, amplification, and

    stabilization phases. The accumulation of neutrophils at the

    injury site was observed to be initiated by a few scouting

    neutrophils. These scouting neutrophils appeared to move

    randomly within the tissue at basal speeds, but upon their

    arrival at the injury foci, additional waves of neutrophils

    rapidly moved through the interstitium with markedly

    increased velocity and directionality toward the injury foci.

    The amplification of the neutrophil numbers persisted for

    approximately 30 min, before numbers stabilized and

    plateaued off. The scouting phase was found to be depen-

    dent on Gai-mediated signaling, suggesting the involve-ment of chemokine receptors, whereas the amplification

    phase depended on cADPR-mediated signaling, which is

    known to be involved in the recognition of certain DAMPs,

    including those mediated by formyl-peptide receptors [85,

    86]. Similarly, in the T. gondii model, amplification phases

    were described to occur after the arrival of pioneer neu-

    trophils [84]. Before swarming occurred, some neutrophils

    were observed to meander randomly across the cluster

    focus, but after the arrival and arrest of pioneer neutrophils,

    those earlier neutrophils were observed returning to con-

    tribute to swarm formation. Swarming behavior may thus

    operate similarly in the context of a pathogen invasion and

    during sterile injury. A role for monocytes in initiating

    swarming behavior was proposed in the ischemic-reperfu-

    sion injury model [45], and thus neutrophil swarming

    behavior may be initiated by various mechanisms and

    potentially involve different signals.

    At the injury foci

    Upon reaching their target site, neutrophils can perform

    several effector functions. Neutrophils are able to secrete the

    contents of their preformed granules, which contain a

    cocktail of antibacterial peptides and enzymes [87]. Granule

    types include azurophilic (primary), specific (secondary),

    and gelatinase (tertiary), in addition to vesicles [88]. Neu-

    trophils can also generate and release toxic reactive oxygen

    species (ROS) through NADPH oxidase in a respiratory

    burst mechanism [89]. Additionally, neutrophils are profi-

    cient phagocytes and are able to hunt down and engulf

    opsonized bacteria and parasites, which include those

    opsonized by either complement or immunoglobulins [49].

    Engulfed contents are then subjected to enzymatic degra-

    dation in the phagolysosomes, which contain additional

    antibacterial peptides and toxic chemicals as a decontami-

    nation measure. More recently, there has been growing

    evidence for an additional method of pathogen defense

    involving the release of neutrophil extracellular traps (NETs)

    [90], which comprise of dense strands of DNA and proteins

    released from within the cell body. The DNA found within

    NETs is believed to be nuclear chromosomal DNA, but some

    studies have shown mitochondrial DNA to be possible as

    well [91]. NETs contain several highly positively charged

    molecules such as histones and are able to trap negatively

    charged bacteria that make contact. In addition to DNA,

    other antibacterial peptides and proteases are extruded,

    including elastase, cathepsins, and lactoferrins among many

    others. Most of these studies about NETs relied on in vitro

    Fig. 3 Schematic showing paracellular and transcellular routes oftransmigration by neutrophils across endothelial wall

    Singapore Immunology Network: SIgN (2012) 53:168181 175

    123

  • approaches. To the best of our knowledge, a study from

    Gunzers laboratory represents the first description of using

    MP imaging to study the formation of NETs in mice in vivo.

    Using the lung slice approach, they were able to image NETs

    and neutrophil behavior in Aspergillus fumigatus-infected

    lungs [92]. The process of forming NETs is not an uncon-

    trolled cell burst and appears to follow a well-defined pro-

    gram. More than one program exists, depending on the

    context of activation, and may include steps such as the

    systematic breakdown of nuclear envelope, modification of

    the DNA for increased toxicity, and packaging of DNA into

    vesicles for release [61]. Conceivably, NETs function most

    effectively within the blood, since shear forces would flush

    circulating bacteria directly into the mesh of DNA for trap-

    ping and killing, whereas the other three defense mecha-

    nisms mentioned earlier are unlikely to be of any use in this

    setting. Indeed, TLR4-activated platelets trigger the forma-

    tion of NETs by neutrophils [93]since platelets require a

    much higher threshold of TLR4 signaling than neutrophils to

    become activated, their activation signifies that a severe

    infection has already occurred within the blood compartment

    [61]. Thus, neutrophils possess a large variety of weapons in

    their arsenal and can employ very different tactics in

    response to a pathogen invasion.

    In addition, neutrophils are also able to secrete a variety of

    cytokines and thus participate in shaping the immune land-

    scape. For example, during T. gondii infection, neutrophils

    engulf the parasite but are unable to perform killing, thereby

    acting as potential hosts for facilitating pathogen spread.

    Interestingly, however, the depletion of neutrophils resulted

    in a worse prognosis [94], and the protective effect of neu-

    trophils was attributed to their shaping of the subsequent

    adaptive immune response via cytokine production [95].

    Other major functions that neutrophils can perform include

    tissue remodeling [84], as well as antigen presentation to

    specialized cells of the adaptive immune system [96].

    On the other hand, the exact roles of neutrophils at sites

    of sterile injury are not well understood. The recruitment of

    neutrophils to sterile injury sites may be an evolutionary

    response to the strong association between pathogen

    invasion and tissue injury. In fact, in addition to triggering

    similar inflammatory responses, many DAMPs can be

    recognized by the same receptors as PAMPs [40]. Thus,

    neutrophil recruitment during sterile conditions may pri-

    marily represent a precautionary measure against patho-

    gens. There could be advantages to the pre-emptive

    recruitment of neutrophils prior to pathogen entry, such as

    the initiation of speedier and more robust responses upon

    pathogen arrival, which would reduce the likelihood of

    successful countermeasures by the pathogen. The presence

    of injury may also signify invisible pathogens that

    neutrophils are unable to detect, which may include hidden

    intracellular viruses. Consequently, a pro-inflammatory

    microenvironment ensues around the injury site, estab-

    lished largely by TNFa and IL-1 [40]. The exact contri-bution made by neutrophils under this condition is

    uncertain. Neutrophils are thought to release granule con-

    tents (which include proteolytic enzymes such as elastase,

    gelatinase B, and collagenase [87]) and phagocytose cel-

    lular debris for tissue remodeling, but the released granule

    contents also possess some pathogen extermination ability.

    Neutrophil depletion has been associated with the accel-

    erated closure of wounds in the epidermis [97], and this

    phenomenon can be attributed to the ability of neutrophils

    to induce keratinocyte differentiation, which slows down

    keratinocyte hyperproliferation [98]. The increased rates of

    keratinocyte differentiation possibly protect against tumor

    formation, while the delayed wound-healing process may

    be necessary to buy time for the adaptive immune system

    to establish an effective response during events for which

    neutrophils have no satisfactory responses against, such as

    viral infections [98]. This delay may also be necessary for

    the neutrophils to buffer enough time for their own clear-

    ance of the infection, since wound closure might be fol-

    lowed by a period of hypoxia where neutrophils may not

    act effectively (e.g., in producing ROS) [98]. Curiously

    though, neutrophil depletion had no apparent effect on the

    rate of wound healing in the dermis [97, 99].

    Neutrophil clearance

    Many of the effector functions of neutrophils result in

    damage to themselvessecreted granules and ROS are

    toxic to the neutrophils, phagocytic neutrophils become

    unable to continue the process above a certain threshold

    [100], and the release of NETs entails the loss of tran-

    scriptional capability necessary for continued cell survival.

    Consequently, in most cases, activated neutrophils have

    short life spans and will undergo apoptosis and clearance

    by tissue macrophages. This uptake of apoptotic neutro-

    phils by macrophages instructs macrophages to undergo a

    gradual switch from inflammatory to anti-inflammatory

    cytokine production and may thus promote the resolution

    of inflammation [101].

    As for the non-activated neutrophils that fulfill their

    designated patrolling lifetimes, although the majority will

    retire in the liver (29%) and spleen (31%), many will return

    to the bone marrow (32%) for clearance [16, 23]. The

    clearance of these neutrophils is mediated by specialized

    macrophages within the various organs [102]. Homing of

    neutrophils to the liver occurs independently of Gai-med-iated signaling. In contrast, for the spleen, half of the

    homing was attributed to a Gai-dependent pathway [16],and thus at least two distinct pathways regulate neutrophil

    clearance, one of which might be chemokine dependent

    176 Singapore Immunology Network: SIgN (2012) 53:168181

    123

  • [102]. On the other hand, neutrophil homing to the bone

    marrow appears almost fully dependent on CXCR4

    CXCL12 interaction. Cell surface expression of CXCR4 on

    neutrophils gradually increases as they age [26, 103] and

    enables neutrophils that are nearing the end of their

    patrolling lifetimes to return to the bone marrow via

    CXCL12 binding. These neutrophils can then transmigrate

    out of the sinusoids back into the extravascular cavities of

    the bone marrow. However, this sinusoid-to-bone-marrow

    transmigration is CD18 independent [104], which is in

    contrast to transendothelial migration in the periphery. The

    adhesion molecules responsible for sinusoid-to-bone-mar-

    row transmigration are as yet unknown, but CD49d

    (VLA-4 subunit a) has been implicated in the bone-mar-row-to-sinusoid transmigration [105]. Upon entry back into

    the bone marrow, neutrophils initiate apoptosis and depend

    on macrophages for their removal. At this stage, the jour-

    ney of a neutrophil would be concluded, but even then, a

    final role awaits themapoptotic neutrophils stimulate the

    bone marrow macrophages that engulf them to produce

    G-CSF [44], which in turn induces the production and

    survival of successive neutrophil populations. Hence,

    neutrophil production rates are pegged to neutrophil utili-

    zation rates, providing a mechanism for the bone marrow

    to regulate total neutrophil numbers. Therefore, an intricate

    feedback system exists, at least partly mediated by the

    neutrophils themselves, to maintain the integrity of the

    immune surveillance program, ensuring their continued

    vigilance against pathogen attack.

    Conclusion

    A fundamental characteristic of the immune system lies in

    the highly dynamic nature of its cellular components. This

    is perhaps best represented by the in vivo behavior of

    neutrophils, which have the ability to rapidly shuttle

    between different body compartments to carry out their

    effector functions in a timely manner. The studies dis-

    cussed in this review have employed different IVM

    approaches to study neutrophil behavior in various tissues,

    uncovering some of the previously unknown or underap-

    preciated aspects of neutrophil functions. Current and

    emerging evidence implies strongly that many of the

    complexities of neutrophil behavior can only be revealed

    when these processes are studied in real-time in vivo and in

    stimulus- and tissue-specific manners. We thus have reason

    to believe that IVM, especially that of MP-IVM, will

    continue to play a vital role in future investigations of the

    cellular functions of neutrophils in vivo.

    Acknowledgments We would like to thank Dr. Jo Keeble for hercritical comments on the manuscript and proofreading.

    References

    1. Semerad CL, Liu F, Gregory AD, Stumpf K, Link DC. G-CSF is

    an essential regulator of neutrophil trafficking from the bone

    marrow to the blood. Immunity. 2002;17(4):41323. Epub

    2002/10/22. PubMed PMID: 12387736.

    2. Dancey JT, Deubelbeiss KA, Harker LA, Finch CA. Neutrophil

    kinetics in man. J Clin Invest. 1976;58(3):70515. Epub

    1976/09/01. doi:10.1172/JCI108517. PubMed PMID: 956397;

    PubMed Central PMCID: PMC333229.

    3. Babior BM, Golde DW. Production, distribution, and fate of

    neutrophils. In: Beutler E, Coller BS, Lichtman MA, Kipps TJ,

    Seligsohn U, editors. Williams Hematology. New York, USA:

    McGraw-Hill; 2001. pp. 7539.

    4. Pillay J, den Braber I, Vrisekoop N, Kwast LM, de Boer RJ,

    Borghans JA, et al. In vivo labeling with 2H2O reveals a human

    neutrophil lifespan of 5.4 days. Blood. 2010;116(4):6257.

    Epub 2010/04/23. doi:10.1182/blood-2010-01-259028. PubMed

    PMID: 20410504.

    5. Ng LG, Qin JS, Roediger B, Wang Y, Jain R, Cavanagh LL,

    et al. Visualizing the neutrophil response to sterile tissue injury

    in mouse dermis reveals a three-phase cascade of events.

    J Invest Dermatol. 2011;131(10):205868. Epub 2011/06/24.

    doi:10.1038/jid.2011.179. PubMed PMID: 21697893.

    6. Waller A. Microscopic examination of some principal tissues of

    the animal frame as observed in the tongue of the living frog,

    toad, etc. Philos Mag. 1846;29:27187.

    7. Germain RN, Miller MJ, Dustin ML, Nussenzweig MC.

    Dynamic imaging of the immune system: progress, pitfalls and

    promise. Nat Rev Immunol. 2006;6(7):497507. Epub 2006/06/

    27. doi:10.1038/nri1884. PubMed PMID: 16799470.

    8. Cahalan MD, Parker I, Wei SH, Miller MJ. Two-photon tissue

    imaging: seeing the immune system in a fresh light. Nat Rev

    Immunol. 2002;2(11):87280. Epub 2002/11/05. doi:10.1038/

    nri935. PubMed PMID: 12415310; PubMed Central PMCID:

    PMC2749751.

    9. Miller MJ, Wei SH, Parker I, Cahalan MD. Two-photon imaging

    of lymphocyte motility and antigen response in intact lymph

    node. Science. 2002;296(5574):186973. Epub 2002/05/23.

    doi:10.1126/science.1070051. PubMed PMID: 12016203.

    10. Mempel TR, Henrickson SE, Von Andrian UH. T-cell priming

    by dendritic cells in lymph nodes occurs in three distinct phases.

    Nature. 2004;427(6970):1549. Epub 2004/01/09. doi:10.1038/

    nature02238. PubMed PMID: 14712275.

    11. Bousso P, Robey E. Dynamics of CD8?T cell priming by

    dendritic cells in intact lymph nodes. Nat Immunol. 2003;

    4(6):57985. Epub 2003/05/06. doi:10.1038/ni928. PubMed

    PMID: 12730692.

    12. Bousso P, Bhakta NR, Lewis RS, Robey E. Dynamics of thy-

    mocyte-stromal cell interactions visualized by two-photon

    microscopy. Science. 2002;296(5574):187680. Epub 2002/06/

    08. doi:10.1126/science.1070945. PubMed PMID: 12052962.

    13. Cahalan MD, Parker I. Choreography of cell motility and

    interaction dynamics imaged by two-photon microscopy in

    lymphoid organs. Annu Rev Immunol. 2008;26:585-626. Epub

    2008/01/05. doi:10.1146/annurev.immunol.24.021605.090620.

    PubMed PMID: 18173372; PubMed Central PMCID: PMC27

    32400.

    14. Ng LG, Mrass P, Kinjyo I, Reiner SL, Weninger W. Two-photon

    imaging of effector T-cell behavior: lessons from a tumor

    model. Immunol Rev. 2008;221:14762. Epub 2008/02/16.

    doi:10.1111/j.1600-065X.2008.00596.x. PubMed PMID: 18275

    480.

    15. Mueller SN, Hickman HD. In vivo imaging of the T cell

    response to infection. Curr Opin Immunol. 2010;22(3):2938.

    Singapore Immunology Network: SIgN (2012) 53:168181 177

    123

  • Epub 2010/01/19. doi:10.1016/j.coi.2009.12.009. PubMed PMID:

    20080040.

    16. Furze RC, Rankin SM. The role of the bone marrow in neu-

    trophil clearance under homeostatic conditions in the mouse.

    Faseb J. 2008;22(9):31119. Epub 2008/05/30. doi:10.1096/

    fj.08-109876. PubMed PMID: 18509199; PubMed Central

    PMCID: PMC2593561.

    17. Dale DC, Liles WC. Neutrophils and monocytes: Normal

    physiology and disorders of neutrophil and monocyte produc-

    tion. In: Handin RI, Lux SE, Stossel TP, editors. Blood:

    principles and practice of hematology. Philadelphia, USA:

    Lippincott Williams & Wilkins; 2003.

    18. Orkin SH, Zon LI. Hematopoiesis: an evolving paradigm for

    stem cell biology. Cell. 2008;132(4):63144. Epub 2008/02/26.

    doi:10.1016/j.cell.2008.01.025. PubMed PMID: 18295580; Pub-

    Med Central PMCID: PMC2628169.

    19. Boxio R, Bossenmeyer-Pourie C, Steinckwich N, Dournon C,

    Nusse O. Mouse bone marrow contains large numbers of func-

    tionally competent neutrophils. J Leukoc Biol. 2004;75(4):

    60411. Epub 2003/12/25. doi:10.1189/jlb.0703340. PubMed

    PMID: 14694182.

    20. Ashtekar AR, Saha B. Polys plea: membership to the club of

    APCs. Trends Immunol. 2003;24(9):48590. doi:10.1016/s1471-

    4906(03)00235-7.

    21. Sabroe I, Prince LR, Jones EC, Horsburgh MJ, Foster SJ, Vogel

    SN, et al. Selective roles for Toll-like receptor (TLR)2 and

    TLR4 in the regulation of neutrophil activation and life span.

    J Immunol. 2003;170(10):526875. Epub 2003/05/08. PubMed

    PMID: 12734376.

    22. John B, Hunter CA. Immunology. Neutrophil soldiers or Trojan

    Horses? Science. 2008;321(5891):9178. Epub 2008/08/16.

    doi:10.1126/science.1162914. PubMed PMID: 18703727.

    23. Suratt BT, Young SK, Lieber J, Nick JA, Henson PM, Worthen

    GS. Neutrophil maturation and activation determine anatomic

    site of clearance from circulation. Am J Physiol Lung Cell Mol

    Physiol. 2001;281(4):L91321. Epub 2001/09/15. PubMed

    PMID: 11557595.

    24. Athens JW, Haab OP, Raab SO, Mauer AM, Ashenbrucker H,

    Cartwright GE, et al. Leukokinetic studies. IV. The total blood,

    circulating and marginal granulocyte pools and the granulocyte

    turnover rate in normal subjects. J Clin Invest. 1961;40:98995.

    Epub 1961/06/01. doi:10.1172/JCI104338. PubMed PMID:

    13684958; PubMed Central PMCID: PMC290816.

    25. Kohler A, De Filippo K, Hasenberg M, van den Brandt C, Nye

    E, Hosking MP, et al. G-CSF-mediated thrombopoietin release

    triggers neutrophil motility and mobilization from bone marrow

    via induction of Cxcr2 ligands. Blood. 2011;117(16):434957.

    Epub 2011/01/13. doi:10.1182/blood-2010-09-308387. PubMed

    PMID: 21224471; PubMed Central PMCID: PMC3087483.

    26. Martin C, Burdon PC, Bridger G, Gutierrez-Ramos JC, Williams

    TJ, Rankin SM. Chemokines acting via CXCR2 and CXCR4

    control the release of neutrophils from the bone marrow and

    their return following senescence. Immunity. 2003;19(4):58393.

    Epub 2003/10/18. PubMed PMID: 14563322.

    27. Bleul CC, Farzan M, Choe H, Parolin C, Clark-Lewis I, So-

    droski J, et al. The lymphocyte chemoattractant SDF-1 is a

    ligand for LESTR/fusin and blocks HIV-1 entry. Nature.

    1996;382(6594):82933. Epub 1996/08/29. doi:10.1038/

    382829a0. PubMed PMID: 8752280.

    28. Sugiyama T, Kohara H, Noda M, Nagasawa T. Maintenance of

    the hematopoietic stem cell pool by CXCL12-CXCR4 chemo-

    kine signaling in bone marrow stromal cell niches. Immunity.

    2006;25(6):97788. Epub 2006/12/19. doi:10.1016/j.immuni.

    2006.10.016. PubMed PMID: 17174120.

    29. Bozic CR, Gerard NP, von Uexkull-Guldenband C, Kolakowski

    LF, Jr., Conklyn MJ, Breslow R, et al. The murine interleukin 8

    type B receptor homologue and its ligands. Expression and

    biological characterization. J Biol Chem. 1994;269(47):

    293558. Epub 1994/11/25. PubMed PMID: 7961909.

    30. Eash KJ, Means JM, White DW, Link DC. CXCR4 is a key

    regulator of neutrophil release from the bone marrow under

    basal and stress granulopoiesis conditions. Blood. 2009;

    113(19):47119. Epub 2009/03/07. doi:10.1182/blood-2008-09-

    177287. PubMed PMID: 19264920; PubMed Central PMCID:

    PMC2680371.

    31. Suratt BT, Petty JM, Young SK, Malcolm KC, Lieber JG, Nick

    JA, et al. Role of the CXCR4/SDF-1 chemokine axis in circu-

    lating neutrophil homeostasis. Blood. 2004;104(2):56571.

    Epub 2004/04/01. doi:10.1182/blood-2003-10-3638. PubMed

    PMID: 15054039.

    32. Delgado MB, Clark-Lewis I, Loetscher P, Langen H, Thelen M,

    Baggiolini M, et al. Rapid inactivation of stromal cell-derived

    factor-1 by cathepsin G associated with lymphocytes. Eur J

    Immunol. 2001;31(3):699707. Epub 2001/03/10. doi:10.1002/

    1521-4141(200103)31:33.0.CO;2-6. PubMed PMID: 11241273.

    33. Levesque JP, Liu F, Simmons PJ, Betsuyaku T, Senior RM,

    Pham C, et al. Characterization of hematopoietic progenitor

    mobilization in protease-deficient mice. Blood. 2004;104(1):

    6572. Epub 2004/03/11. doi:10.1182/blood-2003-05-1589.

    PubMed PMID: 15010367.

    34. Kim HK, De La Luz Sierra M, Williams CK, Gulino AV, Tosato G.

    G-CSF down-regulation of CXCR4 expression identified as a

    mechanism for mobilization of myeloid cells. Blood. 2006;

    108(3):812-20. Epub 2006/03/16. doi:10.1182/blood-2005-10-

    4162. PubMed PMID: 16537807; PubMed Central PMCID:

    PMC1895847.

    35. Delano MJ, Kelly-Scumpia KM, Thayer TC, Winfield RD,

    Scumpia PO, Cuenca AG, et al. Neutrophil mobilization from the

    bone marrow during polymicrobial sepsis is dependent on CXCL12

    signaling. J Immunol. 2011;187(2):9118. Epub 2011/06/22.

    doi:10.4049/jimmunol.1100588. PubMed PMID: 21690321.

    36. Junt T, Schulze H, Chen Z, Massberg S, Goerge T, Krueger A,

    et al. Dynamic visualization of thrombopoiesis within bone

    marrow. Science. 2007;317(5845):176770. Epub 2007/09/22.

    doi:10.1126/science.1146304. PubMed PMID: 17885137.

    37. Seong SY, Matzinger P. Hydrophobicity: an ancient damage-

    associated molecular pattern that initiates innate immune

    responses. Nat Rev Immunol. 2004;4(6):46978. Epub 2004/06/

    03. doi:10.1038/nri1372. PubMed PMID: 15173835.

    38. Hayashi F, Means TK, Luster AD. Toll-like receptors stimulate

    human neutrophil function. Blood. 2003;102(7):26609. Epub

    2003/06/28. doi:10.1182/blood-2003-04-1078. PubMed PMID:

    12829592.

    39. Parker LC, Whyte MK, Dower SK, Sabroe I. The expression and

    roles of Toll-like receptors in the biology of the human neu-

    trophil. J Leukoc Biol. 2005;77(6):88692. Epub 2005/02/25.

    doi:10.1189/jlb.1104636. PubMed PMID: 15728244.

    40. Chen GY, Nunez G. Sterile inflammation: sensing and reacting

    to damage. Nat Rev Immunol. 2010;10(12):82637. Epub

    2010/11/23. doi:10.1038/nri2873. PubMed PMID: 21088683;

    PubMed Central PMCID: PMC3114424.

    41. von Vietinghoff S, Ley K. Homeostatic regulation of blood neu-

    trophil counts. J Immunol. 2008;181(8):51838. Epub 2008/10/

    04. PubMed PMID: 18832668; PubMed Central PMCID:

    PMC2745132.

    42. Pachter JS, de Vries HE, Fabry Z. The blood-brain barrier and

    its role in immune privilege in the central nervous system.

    J Neuropathol Exp Neurol. 2003;62(6):593604. Epub 2003/07/

    02. PubMed PMID: 12834104.

    43. Kim JV, Kang SS, Dustin ML, McGavern DB. Myelomonocytic

    cell recruitment causes fatal CNS vascular injury during acute

    178 Singapore Immunology Network: SIgN (2012) 53:168181

    123

  • viral meningitis. Nature. 2009;457(7226):1915. Epub 2008/11/

    18. doi:10.1038/nature07591. PubMed PMID: 19011611; Pub-

    Med Central PMCID: PMC2702264.

    44. Summers C, Rankin SM, Condliffe AM, Singh N, Peters AM,

    Chilvers ER. Neutrophil kinetics in health and disease. Trends

    Immunol. 2010;31(8):31824. Epub 2010/07/14. doi:10.1016/

    j.it.2010.05.006. PubMed PMID: 20620114; PubMed Central

    PMCID: PMC2930213.

    45. Kreisel D, Nava RG, Li W, Zinselmeyer BH, Wang B, Lai J,

    et al. In vivo two-photon imaging reveals monocyte-dependent

    neutrophil extravasation during pulmonary inflammation. Proc

    Natl Acad Sci USA. 2010;107(42):180738. Epub 2010/10/07.

    doi:10.1073/pnas.1008737107. PubMed PMID: 20923880;

    PubMed Central PMCID: PMC2964224.

    46. Mitchell GF, Jeron A, Koren G. Measurement of heart rate and

    Q-T interval in the conscious mouse. Am J Physiol. 1998;274

    (3 Pt 2):H74751. Epub 1998/04/08. PubMed PMID: 9530184.

    47. Kamoun WS, Chae SS, Lacorre DA, Tyrrell JA, Mitre M,

    Gillissen MA, et al. Simultaneous measurement of RBC

    velocity, flux, hematocrit and shear rate in vascular networks.

    Nat Methods. 2010;7(8):65560. Epub 2010/06/29. doi:10.1038/

    nmeth.1475. PubMed PMID: 20581828; PubMed Central

    PMCID: PMC2921873.

    48. Mazo IB, Gutierrez-Ramos JC, Frenette PS, Hynes RO, Wagner

    DD, von Andrian UH. Hematopoietic progenitor cell rolling in

    bone marrow microvessels: parallel contributions by endothelial

    selectins and vascular cell adhesion molecule 1. J Exp Med.

    1998;188(3):46574. Epub 1998/08/04. PubMed PMID:

    9687524; PubMed Central PMCID: PMC2212463.

    49. Hickey MJ, Kubes P. Intravascular immunity: the host-pathogen

    encounter in blood vessels. Nat Rev Immunol. 2009;9(5):364

    75. Epub 2009/04/25. doi:10.1038/nri2532. PubMed PMID: 193

    90567.

    50. Weninger W, Ulfman LH, Cheng G, Souchkova N, Quacken-

    bush EJ, Lowe JB, et al. Specialized contributions by alpha(1,3)-

    fucosyltransferase-IV and FucT-VII during leukocyte rolling in

    dermal microvessels. Immunity. 2000;12(6):66576. Epub

    2000/07/14. PubMed PMID: 10894166.

    51. Bohannon J, Cui W, Sherwood E, Toliver-Kinsky T. Dendritic

    cell modification of neutrophil responses to infection after burn

    injury. J Immunol. 2010;185(5):284753. Epub 2010/08/04.

    doi:10.4049/jimmunol.0903619. PubMed PMID: 20679533;

    PubMed Central PMCID: PMC3100157.

    52. Ng LG, Hsu A, Mandell MA, Roediger B, Hoeller C, Mrass P,

    et al. Migratory dermal dendritic cells act as rapid sensors of

    protozoan parasites. PLoS Pathog. 2008;4(11):e1000222. Epub

    2008/12/02. doi:10.1371/journal.ppat.1000222. PubMed PMID:

    19043558; PubMed Central PMCID: PMC2583051.

    53. Malaviya R, Ikeda T, Ross E, Abraham SN. Mast cell modulation

    of neutrophil influx and bacterial clearance at sites of infection

    through TNF-alpha. Nature. 1996;381(6577):7780. Epub

    1996/05/02. doi:10.1038/381077a0. PubMed PMID: 8609993.

    54. Beck-Schimmer B, Schwendener R, Pasch T, Reyes L, Booy C,

    Schimmer RC. Alveolar macrophages regulate neutrophil

    recruitment in endotoxin-induced lung injury. Respir Res.

    2005;6:61. Epub 2005/06/24. doi:10.1186/1465-9921-6-61.

    PubMed PMID: 15972102; PubMed Central PMCID: PMC118

    8075.

    55. Mukai K, Matsuoka K, Taya C, Suzuki H, Yokozeki H,

    Nishioka K, et al. Basophils play a critical role in the devel-

    opment of IgE-mediated chronic allergic inflammation inde-

    pendently of T cells and mast cells. Immunity. 2005;23(2):

    191202. Epub 2005/08/23. doi:10.1016/j.immuni.2005.06.011.

    PubMed PMID: 16111637.

    56. Auffray C, Fogg D, Garfa M, Elain G, Join-Lambert O, Kayal S,

    et al. Monitoring of blood vessels and tissues by a population of

    monocytes with patrolling behavior. Science. 2007;317(5838):

    66670. Epub 2007/08/04. doi:10.1126/science.1142883. Pub-

    Med PMID: 17673663.

    57. McDonald B, Pittman K, Menezes GB, Hirota SA, Slaba I,

    Waterhouse CC, et al. Intravascular danger signals guide neu-

    trophils to sites of sterile inflammation. Science. 2010;330

    (6002):3626. Epub 2010/10/16. doi:10.1126/science.1195491.

    PubMed PMID: 20947763.

    58. Chappell D, Westphal M, Jacob M. The impact of the glycocalyx on

    microcirculatory oxygen distribution in critical illness. Curr Opin

    Anaesthesiol. 2009;22(2):15562. Epub 2009/03/25. doi:10.1097/

    ACO.0b013e328328d1b6. PubMed PMID: 19307890.

    59. Massena S, Christoffersson G, Hjertstrom E, Zcharia E,

    Vlodavsky I, Ausmees N, et al. A chemotactic gradient seques-

    tered on endothelial heparan sulfate induces directional intra-

    luminal crawling of neutrophils. Blood. 2010;116(11):192431.

    Epub 2010/06/10. doi:10.1182/blood-2010-01-266072. PubMed

    PMID: 20530797; PubMed Central PMCID: PMC3173988.

    60. Zhang Q, Raoof M, Chen Y, Sumi Y, Sursal T, Junger W, et al.

    Circulating mitochondrial DAMPs cause inflammatory respon-

    ses to injury. Nature. 2010;464(7285):1047. Epub 2010/03/06.

    doi:10.1038/nature08780. PubMed PMID: 20203610; PubMed

    Central PMCID: PMC2843437.

    61. Phillipson M, Kubes P. The neutrophil in vascular inflammation.

    Nat Med. 2011;17(11):138190. Epub 2011/11/09. doi 10.1038/

    nm.2514. PubMed PMID: 22064428.

    62. Heit B, Robbins SM, Downey CM, Guan Z, Colarusso P, Miller

    BJ, et al. PTEN functions to prioritize chemotactic cues and

    prevent distraction in migrating neutrophils. Nat Immunol.

    2008;9(7):74352. Epub 2008/06/10. doi:10.1038/ni.1623. Pub-

    Med PMID: 18536720.

    63. Foxman EF, Campbell JJ, Butcher EC. Multistep navigation and

    the combinatorial control of leukocyte chemotaxis. J Cell Biol.

    1997;139(5):134960. Epub 1998/01/07. PubMed PMID: 938

    2879; PubMed Central PMCID: PMC2140208.

    64. Sogawa Y, Ohyama T, Maeda H, Hirahara K. Inhibition of

    neutrophil migration in mice by mouse formyl peptide receptors

    1 and 2 dual agonist: indication of cross-desensitization in vivo.

    Immunology. 2011;132(3):44150. Epub 2010/11/03. doi:10.

    1111/j.1365-2567.2010.03367.x. PubMed PMID: 21039475;

    PubMed Central PMCID: PMC3044910.

    65. Niethammer P, Grabher C, Look AT, Mitchison TJ. A tissue-

    scale gradient of hydrogen peroxide mediates rapid wound

    detection in zebrafish. Nature. 2009;459(7249):9969. Epub

    2009/06/06. doi:10.1038/nature08119. PubMed PMID: 19494

    811; PubMed Central PMCID: PMC2803098.

    66. Ley K, Laudanna C, Cybulsky MI, Nourshargh S. Getting to the

    site of inflammation: the leukocyte adhesion cascade updated.

    Nat Rev Immunol. 2007;7(9):67889. Epub 2007/08/25.

    doi:10.1038/nri2156. PubMed PMID: 17717539.

    67. Coxon A, Cullere X, Knight S, Sethi S, Wakelin MW, Stavrakis G,

    et al. Fc gamma RIII mediates neutrophil recruitment to immune

    complexes. a mechanism for neutrophil accumulation in immune-

    mediated inflammation. Immunity. 2001;14(6):693704. Epub

    2001/06/23. PubMed PMID: 11420040.

    68. Hentzen ER, Neelamegham S, Kansas GS, Benanti JA, McIntire

    LV, Smith CW, et al. Sequential binding of CD11a/CD18 and

    CD11b/CD18 defines neutrophil capture and stable adhesion to

    intercellular adhesion molecule-1. Blood. 2000;95(3):91120.

    Epub 2000/01/29. PubMed PMID: 10648403.

    69. Phillipson M, Heit B, Colarusso P, Liu L, Ballantyne CM,

    Kubes P. Intraluminal crawling of neutrophils to emigration

    sites: a molecularly distinct process from adhesion in the

    recruitment cascade. J Exp Med. 2006;203(12):256975. Epub

    2006/11/23. doi:10.1084/jem.20060925. PubMed PMID:

    17116736; PubMed Central PMCID: PMC2118150.

    Singapore Immunology Network: SIgN (2012) 53:168181 179

    123

  • 70. Schenkel AR, Mamdouh Z, Muller WA. Locomotion of

    monocytes on endothelium is a critical step during extravasa-

    tion. Nat Immunol. 2004;5(4):393400. Epub 2004/03/17.

    doi:10.1038/ni1051. PubMed PMID: 15021878.

    71. Menezes GB, Lee WY, Zhou H, Waterhouse CC, Cara DC,

    Kubes P. Selective down-regulation of neutrophil Mac-1 in

    endotoxemic hepatic microcirculation via IL-10. J Immunol.

    2009;183(11):755768. Epub 2009/11/18. doi:10.4049/jimmu

    nol.0901786. PubMed PMID: 19917697.

    72. Feng D, Nagy JA, Pyne K, Dvorak HF, Dvorak AM. Neutrophils

    emigrate from venules by a transendothelial cell pathway in

    response to FMLP. J Exp Med. 1998;187(6):90315. Epub

    1998/04/04. PubMed PMID: 9500793; PubMed Central PMCID:

    PMC2212194.

    73. Phillipson M, Kaur J, Colarusso P, Ballantyne CM, Kubes P.

    Endothelial domes encapsulate adherent neutrophils and mini-

    mize increases in vascular permeability in paracellular and

    transcellular emigration. PLoS One. 2008;3(2):e1649. Epub

    2008/02/26. doi:10.1371/journal.pone.0001649. PubMed PMID:

    18297135; PubMed Central PMCID: PMC2250804.

    74. Woodfin A, Voisin MB, Beyrau M, Colom B, Caille D, Diapouli

    FM, et al. The junctional adhesion molecule JAM-C regulates

    polarized transendothelial migration of neutrophils in vivo. Nat

    Immunol. 2011;12(8):7619. Epub 2011/06/28. doi:10.1038/

    ni.2062. PubMed PMID: 21706006; PubMed Central PMCID:

    PMC3145149.

    75. Phillipson M, Heit B, Parsons SA, Petri B, Mullaly SC,

    Colarusso P, et al. Vav1 is essential for mechanotactic crawling

    and migration of neutrophils out of the inflamed micro-

    vasculature. J Immunol. 2009;182(11):68708. Epub 2009/05/

    21. doi:10.4049/jimmunol.0803414. PubMed PMID: 19454683.

    76. Burns AR, Walker DC, Brown ES, Thurmon LT, Bowden RA,

    Keese CR, et al. Neutrophil transendothelial migration is inde-

    pendent of tight junctions and occurs preferentially at tricellular

    corners. J Immunol. 1997;159(6):2893903. Epub 1997/09/23.

    PubMed PMID: 9300713.

    77. Schmidt EP, Lee WL, Zemans RL, Yamashita C, Downey GP.

    On, around, and through: neutrophil-endothelial interactions in

    innate immunity. Physiology (Bethesda). 2011;26(5):33447.

    Epub 2011/10/21. doi:10.1152/physiol.00011.2011. PubMed

    PMID: 22013192.

    78. Barreiro O, Zamai M, Yanez-Mo M, Tejera E, Lopez-Romero P,

    Monk PN, et al. Endothelial adhesion receptors are recruited to

    adherent leukocytes by inclusion in preformed tetraspanin

    nanoplatforms. J Cell Biol. 2008;183(3):52742. Epub 2008/10/

    29. doi:10.1083/jcb.200805076. PubMed PMID: 18955551;

    PubMed Central PMCID: PMC2575792.

    79. Barreiro O, Yanez-Mo M, Serrador JM, Montoya MC, Vicente-

    Manzanares M, Tejedor R, et al. Dynamic interaction of

    VCAM-1 and ICAM-1 with moesin and ezrin in a novel endo-

    thelial docking structure for adherent leukocytes. J Cell Biol.

    2002;157(7):123345. Epub 2002/06/26. doi:10.1083/jcb.20011

    2126. PubMed PMID: 12082081; PubMed Central PMCID:

    PMC2173557.

    80. Muller WA. Leukocyte-endothelial-cell interactions in leuko-

    cyte transmigration and the inflammatory response. Trends

    Immunol. 2003;24(6):32734. Epub 2003/06/18. PubMed

    PMID: 12810109.

    81. Carman CV, Sage PT, Sciuto TE, de la Fuente MA, Geha RS,

    Ochs HD, et al. Transcellular diapedesis is initiated by invasive

    podosomes. Immunity. 2007;26(6):78497. Epub 2007/06/16.

    doi:10.1016/j.immuni.2007.04.015. PubMed PMID: 17570692;

    PubMed Central PMCID: PMC2094044.

    82. Carman CV. Mechanisms for transcellular diapedesis: probing

    and pathfinding by invadosome-like protrusions. J Cell Sci.

    2009;122(Pt 17):302535. Epub 2009/08/21. doi:10.1242/

    jcs.047522. PubMed PMID: 19692589.

    83. Carman CV, Springer TA. A transmigratory cup in leukocyte

    diapedesis both through individual vascular endothelial cells and

    between them. J Cell Biol. 2004;167(2):37788. Epub 2004/10/

    27. doi:10.1083/jcb.200404129. PubMed PMID: 15504916;

    PubMed Central PMCID: PMC2172560.

    84. Chtanova T, Schaeffer M, Han SJ, van Dooren GG, Nollmann

    M, Herzmark P, et al. Dynamics of neutrophil migration in

    lymph nodes during infection. Immunity. 2008;29(3):48796.

    Epub 2008/08/23. doi:10.1016/j.immuni.2008.07.012. PubMed

    PMID: 18718768; PubMed Central PMCID: PMC2569002.

    85. Partida-Sanchez S, Cockayne DA, Monard S, Jacobson EL,

    Oppenheimer N, Garvy B, et al. Cyclic ADP-ribose production

    by CD38 regulates intracellular calcium release, extracellular

    calcium influx and chemotaxis in neutrophils and is required for

    bacterial clearance in vivo. Nat Med. 2001;7(11):120916. Epub

    2001/11/02. doi:10.1038/nm1101-1209. PubMed PMID: 116

    89885.

    86. Partida-Sanchez S, Iribarren P, Moreno-Garcia ME, Gao JL,

    Murphy PM, Oppenheimer N, et al. Chemotaxis and calcium

    responses of phagocytes to formyl peptide receptor ligands is

    differentially regulated by cyclic ADP ribose. J Immunol.

    2004;172(3):1896906. Epub 2004/01/22. PubMed PMID:

    14734775.

    87. Borregaard N, Cowland JB. Granules of the human neutrophilic

    polymorphonuclear leukocyte. Blood. 1997;89(10):350321.

    Epub 1997/05/15. PubMed PMID: 9160655.

    88. Lacy P. Mechanisms of degranulation in neutrophils. Allergy

    Asthma Clin Immunol. 2006;2(3):98108. Epub 2006/09/15.

    doi:10.1186/1710-1492-2-3-98. PubMed PMID: 20525154;

    PubMed Central PMCID: PMC2876182.

    89. de Oliveira-Junior EB, Bustamante J, Newburger PE, Condino-

    Neto A. The human NADPH oxidase: primary and secondary

    defects impairing the respiratory burst function and the micro-

    bicidal ability of phagocytes. Scand J Immunol. 2011;73(5):

    4207. Epub 2011/01/06. doi:10.1111/j.1365-3083.2010.025

    01.x. PubMed PMID: 21204900.

    90. Brinkmann V, Reichard U, Goosmann C, Fauler B, Uhlemann

    Y, Weiss DS, et al. Neutrophil extracellular traps kill bacteria.

    Science. 2004;303(5663):15325. Epub 2004/03/06. doi:10.

    1126/science.1092385. PubMed PMID: 15001782.

    91. Yousefi S, Mihalache C, Kozlowski E, Schmid I, Simon HU.

    Viable neutrophils release mitochondrial DNA to form neutro-

    phil extracellular traps. Cell Death Differ. 2009;16(11):

    143844. Epub 2009/07/18. doi:10.1038/cdd.2009.96. PubMed

    PMID: 19609275.

    92. Bruns S, Kniemeyer O, Hasenberg M, Aimanianda V, Nietzsche

    S, Thywissen A, et al. Production of extracellular traps against

    Aspergillus fumigatus in vitro and in infected lung tissue is

    dependent on invading neutrophils and influenced by

    hydrophobin RodA. PLoS Pathog. 2010;6(4):e1000873. Epub

    2010/05/06. doi:10.1371/journal.ppat.1000873. PubMed PMID:

    20442864; PubMed Central PMCID: PMC2861696.

    93. Clark SR, Ma AC, Tavener SA, McDonald B, Goodarzi Z, Kelly

    MM, et al. Platelet TLR4 activates neutrophil extracellular traps

    to ensnare bacteria in septic blood. Nat Med. 2007;13(4):4639.

    Epub 2007/03/27. doi:10.1038/nm1565. PubMed PMID: 173

    84648.

    94. Bliss SK, Gavrilescu LC, Alcaraz A, Denkers EY. Neutrophil

    depletion during Toxoplasma gondii infection leads to impaired

    immunity and lethal systemic pathology. Infect Immun. 2001;

    69(8):4898905. Epub 2001/07/12. doi:10.1128/IAI.69.8.4898-

    4905.2001. PubMed PMID: 11447166; PubMed Central PMCID:

    PMC98580.

    180 Singapore Immunology Network: SIgN (2012) 53:168181

    123

  • 95. Denkers EY, Butcher BA, Del Rio L, Bennouna S. Neutrophils,

    dendritic cells and Toxoplasma. Int J Parasitol. 2004;34(3):

    41121. Epub 2004/03/09. doi:10.1016/j.ijpara.2003.11.001. Pub-

    Med PMID: 15003500.

    96. Shen L, Rock KL. Priming of T cells by exogenous antigen

    cross-presented on MHC class I molecules. Curr Opin Immunol.

    2006;18(1):8591. Epub 2005/12/06. doi:10.1016/j.coi.2005.11.

    003. PubMed PMID: 16326087.

    97. Dovi JV, He LK, DiPietro LA. Accelerated wound closure in

    neutrophil-depleted mice. J Leukoc Biol. 2003;73(4):44855.

    Epub 2003/03/28. PubMed PMID: 12660219.

    98. Dovi JV, Szpaderska AM, DiPietro LA. Neutrophil function in

    the healing wound: adding insult to injury? Thrombosis and

    Haemostasis. 2004. doi:10.1160/th03-11-0720.

    99. Simpson DM, Ross R. The neutrophilic leukocyte in wound

    repair a study with antineutrophil serum. J Clin Invest.

    1972;51(8):200923. Epub 1972/08/01. doi:10.1172/JCI107007.

    PubMed PMID: 5054460; PubMed Central PMCID: PMC292

    357.

    100. Simon SI, Schmid-Schonbein GW. Biophysical aspects of

    microsphere engulfment by human neutrophils. Biophys J. 1988;

    53(2):16373. Epub 1988/02/01. doi:10.1016/S0006-3495(88)

    83078-9. PubMed PMID: 3345329; PubMed Central PMCID:

    PMC1330137.

    101. Fadok VA, McDonald PP, Bratton DL, Henson PM. Regulation

    of macrophage cytokine production by phagocytosis of apop-

    totic and post-apoptotic cells. Biochem Soc Trans. 1998;

    26(4):6536. Epub 1999/02/27. PubMed PMID: 10047800.

    102. Rankin SM. The bone marrow: a site of neutrophil clearance.

    J Leukoc Biol. 2010;88(2):24151. Epub 2010/05/21. doi:10.

    1189/jlb.0210112. PubMed PMID: 20483920.

    103. Nagase H, Miyamasu M, Yamaguchi M, Imanishi M, Tsuno

    NH, Matsushima K, et al. Cytokine-mediated regulation of

    CXCR4 expression in human neutrophils. J Leukoc Biol. 2002;

    71(4):7117. Epub 2002/04/03. PubMed PMID: 11927659.

    104. Stark MA, Huo Y, Burcin TL, Morris MA, Olson TS, Ley K.

    Phagocytosis of apoptotic neutrophils regulates granulopoiesis

    via IL-23 and IL-17. Immunity. 2005;22(3):28594. Epub

    2005/03/23. doi:10.1016/j.immuni.2005.01.011. PubMed PMID:

    15780986.

    105. Burdon PC, Martin C, Rankin SM. The CXC chemokine MIP-2

    stimulates neutrophil mobilization from the rat bone marrow in a

    CD49d-dependent manner. Blood. 2005;105(6):25438. Epub

    2004/11/16. doi:10.1182/blood-2004-08-3193. PubMed PMID:

    15542579.

    Singapore Immunology Network: SIgN (2012) 53:168181 181

    123

    Peeking into the secret life of neutrophilsPeeking into the secret life of neutrophilsIntroductionThe behavior of neutrophils in different tissue compartmentsBone marrowBloodBlood vessel wallsInterstitiumAt the injury foci

    Neutrophil clearanceConclusionAcknowledgmentsReferences