Next Article in Journal
Pharmacological Characterization of P626, a Novel Dual Adenosine A2A/A2B Receptor Antagonist, on Synaptic Plasticity and during an Ischemic-like Insult in CA1 Rat Hippocampus
Next Article in Special Issue
Does SARS-CoV-2 Induce IgG4 Synthesis to Evade the Immune System?
Previous Article in Journal
The Functional Role of Group 2 Innate Lymphoid Cells in Asthma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Targeting Neutrophil β2-Integrins: A Review of Relevant Resources, Tools, and Methods

by
Haleigh E. Conley
1,2 and
M. Katie Sheats
1,2,*
1
Department of Clinical Sciences, College of Veterinary Medicine, North Carolina State University, Raleigh, NC 27607, USA
2
Comparative Medicine Institute, North Carolina State University, Raleigh, NC 27607, USA
*
Author to whom correspondence should be addressed.
Biomolecules 2023, 13(6), 892; https://doi.org/10.3390/biom13060892
Submission received: 26 April 2023 / Revised: 15 May 2023 / Accepted: 23 May 2023 / Published: 26 May 2023
(This article belongs to the Special Issue Insights of Innate Immunology into Inflammation and Infections)

Abstract

:
Neutrophils are important innate immune cells that respond during inflammation and infection. These migratory cells utilize β2-integrin cell surface receptors to move out of the vasculature into inflamed tissues and to perform various anti-inflammatory responses. Although critical for fighting off infection, neutrophil responses can also become dysregulated and contribute to disease pathophysiology. In order to limit neutrophil-mediated damage, investigators have focused on β2-integrins as potential therapeutic targets, but so far these strategies have failed in clinical trials. As the field continues to move forward, a better understanding of β2-integrin function and signaling will aid the design of future therapeutics. Here, we provide a detailed review of resources, tools, experimental methods, and in vivo models that have been and will continue to be utilized to investigate the vitally important cell surface receptors, neutrophil β2-integrins.

1. Introduction

Neutrophils are the predominant circulating leukocytes in the blood and are considered the first responders of the immune system. Neutrophils defend the host against invading pathogens via effector functions such as respiratory burst, phagocytosis, and the release of NETs (see Abbreviations). To accomplish these tasks, neutrophils must travel out of the vasculature and into the injured or infected tissue through a process known as transmigration. β2-integrins are specialized cell surface receptors that play a key role in a neutrophil’s ability to transmigrate. The mechanisms of β2-integrin function and signaling have been researched and reviewed extensively [1,2]. While much has been learned, therapeutic efforts to target β2-integrins to mitigate neutrophil-mediated host injury or disease have not proved clinically beneficial. Because neutrophils play a role in the pathophysiology of diseases ranging from acute lung injury and sepsis to rheumatoid arthritis and organ transplant rejection, the methods used to study neutrophil β2-integrins, reviewed here, remain of interest to a wide array of basic, translational, and clinical health researchers.

2. β2-Integrins and Neutrophils

2.1. β2-Integrin Activation and Signaling

Expressed exclusively on leukocytes, β2-integrins are transmembrane heterodimers that consist of a common β-subunit (CD18), which is non-covalently associated with one of the four known α-subunits (CD11a,b,c,d) (Table 1) [3,4]. The two most prominent and most studied integrins on neutrophils are LFA-1 (αLβ2) and Mac-1 (αMβ2). The αDβ2 integrin is the least researched β2-integrin; however, a recent review extensively covers what is known about this integrin [4]. Within circulating, quiescent neutrophils, αMβ2 integrins are primarily contained within the cytoplasm, the secondary and tertiary granules, and the secretory vesicles. On the surface of resting neutrophils, αLβ2-integrins are in an inactive or ‘bent’ conformation, termed the “low-affinity” state. This combination of low surface expression and inactive conformation are control measures to help prevent non-specific neutrophil binding and activation, as unregulated activation could lead to damaging effects for the host. It is only when neutrophils encounter an activation signal, such as the binding of a chemoattractant (e.g., leukotriene B4 (LTB4), N-formylmethionine-leucyl-phenylalanine (fMLP)) to a G-protein-coupled receptor (GPCR), that αMβ2-integrin surface expression is increased, conformational changes take place (“affinity”) to open β2-integrins, and increased mobility within the membrane leads to cluster formation (“avidity/valency”). This method of activation in which integrin affinity and avidity are altered by intracellular signals that affect change at the integrin cytoplasmic tail is known as “inside-out” activation [5,6]. In contrast to “inside-out,” “outside-in” activation occurs when the β2-integrin extracellular domain interacts directly with extracellular matrix proteins or other cell surface ligands (ICAM-1, fibrinogen, etc.) and initiates its own signaling [1,7]. This triggers the phosphorylation of ITAM-bearing transmembrane adapters DAP-12 and Fcγ receptors (FcγRs), which go on to activate Syk and initiate a signaling cascade for cytoskeletal reorganization [8]. These effects on the cytoskeleton are important for the role of β2-integrins in neutrophil adhesion strengthening, cell spreading, and crawling [9]. Despite being described, and studied in vitro, as distinct pathways, inside-out and outside-in activation are designed to work in concert in vivo, with signaling from one pathway reinforcing the other, and vice versa.

2.2. Neutrophils in Disease

Neutrophils are essential as the “first responders” of the immune system, and without these cells patients are at increased risk from infection or injury. The clearest illustration of this is the frequent and life-threatening infections experienced by patients who lack functional β2-integrins due to Leukocyte Adhesion Deficiency (LAD). However, although neutrophils are essential for the maintenance of life and health, they can also cause damage to host tissue in numerous chronic inflammatory conditions and acute inflammatory events, making neutrophil-targeting therapies highly desirable. For example, patients with severe SARS-CoV-2 experience an influx of neutrophils into the lungs, resulting in alveolar damage and the development of acute respiratory distress syndrome (ARDS) [10,11]. Numerous diseases and disorders involve neutrophils and specifically β2-integrins (Table 2). Neutrophil β2-integrins also bind pathogen-associated molecular patterns (PAMPs), such as LPS and β-glucans [12,13,14,15]. Thus, neutrophils can also become activated directly by pathogens during infection, resulting in effector responses, such as respiratory burst, phagocytosis, and NET formation. Because of the essential role integrins play in neutrophil inflammatory functions, they are attractive therapeutic targets [16,17]. However, β2-integrin-targeting therapies have not been successful in clinical trials [18], and additional research is needed to identify new methods of targeting these integrins. Within the scientific literature, there have been many approaches to studying β2-integrins. This review article will provide an overview of methods used for investigating β2-integrin function, activation, and signaling in neutrophils. Hopefully, with the use of resources, tools and methods presented here, and the continued development of new approaches, researchers will discover a more comprehensive understanding of β2-integrins that will lead to successful therapeutic targets to benefit patients with neutrophil-mediated disease.

3. Cell Types and Tools for Evaluating Integrins

3.1. Primary Cells

Neutrophils are hematopoietic cells that are terminally differentiated from myeloblasts. They are the most predominant circulating leukocytes in the blood and their typical life span in circulation ranges between less than 24 h to 5.4 days [56,57]. Most primary neutrophil research uses cells collected from either humans or mice. Over the years, several methods have been utilized to isolate neutrophils from human peripheral blood. Most protocols include erythrocyte sedimentation and density centrifugation. Erythrocyte sedimentation is typically achieved using dextran (varying concentrations 1–6%) or HetaSep [58]. For density centrifugation, Ficoll and Percoll are the most commonly utilized options. The order of these steps often varies depending on the research group. Despite their common usage, there is concern that neutrophils isolated by Dextran and Ficoll are prematurely activated by the presence of monocytes [59]. To avoid this background stimulation, some use a one-step high-density Ficoll (1.114 g/mL) without erythrocyte sedimentation, or substitute a discontinuous gradient for Ficoll [59,60]. Neutrophils may also be “rested” following isolation prior to functional assays to reduce unintended activation [61]. Red blood cell lysis usually follows the isolation of neutrophils; however, when the experimental method does not require the removal of red blood cells, such as flow cytometry, avoiding lysis may be one strategy to prevent unwanted neutrophil activation.
These isolation methods routinely yield normal-density neutrophils (NDNs) but fail to isolate the low-density neutrophils (LDNs) that exist in individuals with inflammatory conditions. To isolate LDNs, negative selection by magnetic beads of both the peripheral blood mononuclear cell (PBMC) layer and the granulocyte layer is necessary [58,62]. While the application of magnetic microbeads facilitates the isolation of pure cell populations, this method increases the cost of isolation and may still require lysis to remove contaminating RBCs [63]. Neutrophils isolated via negative selection by microbeads do not display iatrogenic activation because the labeling antibodies are not directed at neutrophils [64]. In fact, magnetic separation of neutrophils results in significantly lower iatrogenic activation compared with traditional dextran sedimentation followed by density centrifugation with Percoll [65,66].
Primary human neutrophils are easy to obtain from willing donors, and a major benefit of using primary human cells is the ability to obtain samples from humans with diseases of interest. Neutrophils from LAD-I patients have been an invaluable resource for researchers interested in β2-integrin-dependent and independent neutrophil functions and cell signaling downstream of β2-integrins [35,67]. However, the risk to the patient versus the benefit of health discovery research must also be a consideration when obtaining samples from patients. For this reason, volume and cell number are likely to be even more limited when samples are obtained from diseased patients. While sampling from human populations can be extremely convenient, there are logistical considerations, such as Institutional Review Board (IRB) approvals and the need for technically skilled personnel, which often restricts human research with primary neutrophils to experienced labs. This was especially true during the COVID-19 pandemic when IRB protocols changed to reduce the risk for participants and researchers, resulting in reduced access to human participants. Commercially available (e.g., iQ Biosciences®, HemaCare®) cryopreserved products are a potential alternative for researchers seeking to perform experiments with primary neutrophils; however, while cryopreserved neutrophils retain some phagocytic and migratory functions, these are diminished compared with freshly isolated cells [68]. Additionally, preservation of the oxidative metabolism and microbicidal activity requires specialized storage techniques [69].
In addition to primary cells, enucleated neutrophils (cytoplasts) and neutrophil-derived extracellular vesicles also have functional activity. Cytoplasts are enucleated neutrophils that retain very similar chemotactic and bactericidal activity as their parent neutrophils, and these functions are reportedly intact following cryopreservation [70,71]. Neutrophil-derived extracellular vesicles (EVs) also have functional antibacterial activity and express common neutrophil surface receptors, such as CD11b and CD18 [72,73,74]. Interestingly, neutrophil EVs can be stimulated through Mac-1 clustering, highlighting the downstream signaling that occurs following outside-in β2-integrin clustering [75]. Although the information on neutrophil EVs is still limited, future studies should interrogate the intracellular signaling related to the β2-integrin involvement of these EVs to complement the amassing antimicrobial functional data.
Primary neutrophils from mice are commonly isolated from bone marrow or peripheral blood. Similar to humans, density centrifugation with either Percoll or Histopaque discontinuous gradients is utilized [76,77,78]. Due to the wide availability of species-specific resources, murine neutrophils can also be isolated from bulk populations using magnetic microbeads or by fluorescence-activated cell sorting (FACS) [79,80]. Murine neutrophils are widely used due to the availability of models that duplicate neutrophil function during health and disease. Although they express the same β2-integrins, mouse neutrophils do not provide a perfect parallel to humans. In humans, neutrophils are the predominant circulating cell type in the blood (50–70% neutrophils, 30–50% lymphocytes), whereas mice have an abundance of lymphocytes (10–25% neutrophils, 75–90% lymphocytes) [81]. Additionally, murine neutrophils do not express the same FcγRs as human neutrophils [82,83]. Murine neutrophils isolated from bone marrow also display different surface markers and functional activities compared with neutrophils harvested from peripheral blood due to the presence of more immature neutrophils and neutrophil precursors in the bone marrow. Magnetic bead selection is one method that can improve the isolation of mature neutrophils from mouse bone marrow [84,85,86].
Neutrophils are a heterogenous population consisting of normal-density neutrophils (NDNs), low-density neutrophils (LDNs), immature neutrophils, mature neutrophils, and neutrophils with immunosuppressive capabilities [87]. Although mice do express these neutrophil subpopulations, they do not always mirror the presentation of that in humans. During acute infection and inflammation, murine peripheral blood neutrophils exhibit both proinflammatory (CD11bCD49d+IL-12+) and anti-inflammatory (CD11b+CD49dIL-10+) neutrophil subsets that have not yet been identified in humans [87]. Human autoimmune disease often results in increased circulation of proinflammatory LDNs, but murine models of autoimmune disease do not display these heterologous populations of neutrophils [87]. These phenotypic differences have been linked with functional differences as well, specifically in murine cancer models [87,88]. Furthermore, in a study conducted by Soroush et al., stimulation of murine pulmonary endothelial cells with tumor necrosis factor α (TNFα) did not result in the upregulation of ICAM-1 expression whereas TNFα stimulation of human pulmonary endothelial cells did induce ICAM-1 upregulation [89]. Thus, it may be more difficult to model β2-integrin-dependent interactions of neutrophils with murine-derived endothelial cells. Human neutrophils also have significant transcriptional and epigenetic diversity. Females especially have elevated gene expression levels related to immune responses that correspond with increased occurrences of autoimmune disease [90]. Further, mice cannot model the impact of ethnic diversity on neutrophils, despite recent advances in high-diversity mouse populations [91,92].

3.2. Cell Lines

While freshly isolated primary cells are highly desirable for understanding neutrophil β2-integrins, they are not always accessible or suitable for certain experiments. For example, although reported [93,94], the manipulation of RNA and protein expression in primary neutrophils is extremely difficult, so cell lines are beneficial for researchers aiming to investigate the roles of individual proteins through knockdown or overexpression studies. Despite some limitations and drawbacks of neutrophil-like cell lines (Table 3), they can be a useful approach for studying integrin function and signaling via induced mutations, rather than having to develop a new transgenic mouse line.
The HL60 cell line is a human promyeoloblast cell line that can be differentiated into neutrophil-like cells utilizing dimethylsulfoxide (DMSO) or retinoic acid [95,96]. PLB-985 cells are a genetically identical subline of HL60 that are also differentiated using DMSO or retinoic acid [97,98,99]. Both methods of differentiation in HL60s and PLB-985s result in mature neutrophil-like cells; however, compared with DMSO, differentiation with retinoic acid resulted in dampened cellular responses to fMLP and increased random cellular migration [100]. The two methods also result in different expression levels of Scar1 and WASP proteins [101]. Functionally, DMSO-differentiated HL60 neutrophil-like cells are mostly similar to primary neutrophils but do express some differences (Table 3) [95,97,102,103,104]. Despite these differences, research using mutated HL60s has contributed to our understanding of LFA-1 in migration [105]. HL60s have also been used to model host–pathogen interactions [106].
Although not as commonly used as HL60s, the human myeloid cell line K562 has been utilized over the past 15 years in research focusing on granulocytes and β2-integrins. Xue et al. used the K562 cell line to determine the impacts of kindlin-3 defects on integrin function. Through these studies, it was demonstrated that kindlin-3 is required for β2-integrin-mediated adhesion and cell spreading [107]. Another group of researchers used K562 cells as a means to express constructs of αM fused to mCFP and β2 fused to mYFP to assess Mac-1 cytoplasmic tail separation during integrin activation via FRET analysis [108]. With this technique, investigators determined that integrin-ligand binding, or integrin crosslinking, induced Mac-1 cytoplasmic tail separation, which was essential for triggering outside-in signaling pathways. This key finding offers insight into why this strategy of leukocyte adhesion blockade failed in clinical trials, as these integrin ligand mimetics designed to block neutrophil-endothelial adhesion were activating neutrophils through a different pathway [109,110]. K562 cells have been further used to evaluate potential small peptide inhibitors and monoclonal antibodies directed against β2-integrins [111]. Another benefit of K562 cells is that they only express the transfected β2-integrin, allowing for studies examining only Mac-1 or LFA-1, if desired. They also respond similarly to common stimuli of primary neutrophils, including Mn2+ [108,112].
HoxB8 cells are immortalized murine hematopoietic progenitors that are differentiated into neutrophil-like cells using GM-CSF treatment. These cells perform many neutrophil and integrin-mediated functions similar to primary murine neutrophils, with a few discrepancies (Table 3) [113,114,115,116]. The usage of these cells has increased over the past several years, primarily because HoxB8 cells can be generated from transgenic mice and used to examine specific signaling molecule interactions related to β2-integrins [117,118]. They can also be engrafted into naïve mice and functionally respond to bacterial pathogens [116]. Recently, investigators used HoxB8 neutrophil-like cells to show that Rap1 and Riam binding to talin is critical for β2-integrin function [117]. In addition to traditional methods of transfection and siRNA, HoxB8 cells can also be manipulated by CRISPR/Cas9 technology to achieve mutants of interest [117,118]. Importantly, murine HoxB8 neutrophil-like cells expressing a human β2-integrin ortholog display fully functional signaling and adhesive properties in response to common stimuli, such as PMA, TNFα, etc. [118]

3.3. Tools

3.3.1. Anti-Integrin and Fluorescently Labeled Antibodies

Antibodies, including anti-integrin antibodies, are a common tool used to investigate β2-integrins [67,106,119,120,121] (Table 4). These antibodies are advantageous for common lab use due to their ease of application and relatively low cost. They are routinely used in three different ways: function blocking, integrin crosslinking, or fluorescent labeling. As a tool to block function, anti-integrin antibodies led to the discovery that Mac-1 is responsible for neutrophil firm adhesion [67]. However, antibody binding of β2-integrins can also cause activation, as demonstrated by antibody stimulation of neutrophils in the absence of a ligand (e.g., ICAM-1) and subsequent β2-integrin outside-in activation and signaling [108]. This dual nature requires careful experimental planning to prevent unintentional crosslinking and/or Fc receptor engagement when used for inhibitory applications [122]. To avoid these unintentional interactions, researchers can use F(ab) and F(ab)’2 fragments derived from monoclonal antibodies [123,124]. These fragments are portions of antibodies where the Fc fragments are cleaved off to prevent non-specific binding of Fc receptors to antibodies. Both types of fragments are helpful in blocking antibodies, and F(ab)’2 provide additional capabilities for precipitating proteins of interest.
The use of fluorescently labeled antibodies has also been an invaluable tool for researchers. Using specific antibodies, surface expression levels of integrins, including the bent versus open conformations, can be examined quantitatively and qualitatively [125]. This methodology led to the discovery that neutrophils from patients with antiphospholipid syndrome (APS) have an upregulation of activated CD11b, which contributes to increased neutrophil adhesiveness [126]. Fluorescently labeled antibodies can also be used in vivo. Wilson et al. administered PE-anti-Ly6G intravenously to evaluate neutrophil infiltration induced by P. aeruginosa in talin-1 or kindlin-3 knockout mice [127]. As researchers continue to seek novel protein targets to regulate β2-integrins, fluorescently labeled antibodies combined with flow cytometry and/or microscopy may be a first step to understanding the impact inhibitors may have on β2-integrin expression and activation.
Table 4. Common antibodies used to interrogate β2-integrins.
Table 4. Common antibodies used to interrogate β2-integrins.
AntibodyCloneConformation/PurposeReferences
Anti-CD18IB4Recognizes CD18 expression[128,129]
Crosslinking of CD18
In vitro blocking of human β2-integrins
GAME-46Recognizes murine CD18 expression
In vitro and in vivo blocking of murine CD18
[127,130,131]
CBR LFA-1/2Crosslinking of CD18
Recognizes CD18 expression
[105,132]
Anti-CD11bCBMR1/5Recognizes high-affinity /activated CD11b[133,134,135]
Anti-CD11bICRF44Recognizes CD11b expression[136]
M1/70Recognizes CD11b expression
In vivo blocking of murine CD11b
[137,138]
Anti-human β2-integrinKIM127Recognizes bent low-affinity (E+H) β2-integrin conformation[135,139]
Anti-human CD11a/CD18m24Recognizes extended/high-affinity (H+) β2-integrin conformation[139]

3.3.2. Divalent Cations

Divalent cations (e.g., Mn2+, Ca2+, Mg2+) are required for many biological processes, including the binding of integrins to their ligands. Both manganese (Mn2+) and magnesium (Mg2+) act by binding the metal-ion-dependent adhesion site (MIDAS) domain. Mn2+ binding to the MIDAS domain induces outside-in β2-integrin activation by forcing integrins to assume a high-affinity conformation that enhances ligand binding [105]. Because of this effect on the MIDAS domain, Mn2+ can also be applied as a rescue strategy when examining integrin defects caused by mutation or chemical inhibition. In the absence of inside-out activation signals, Mn2+-stimulation can be used as the proximal-most event in the outside-in β2-integrin signaling cascade. Using this approach, Xu et al. determined that Mn2+ treatment could not rescue the binding defects of myosin light chain kinase (MYLK)-deficient murine neutrophils, indicating a critical role for MYLK in outside-in β2-integrin activation [45].
Like manganese, calcium and magnesium are required for biological processes. Therefore, researchers commonly include calcium and magnesium supplementation in media, and manipulation of cation presence has led to a better understanding of integrin regulation. Calcium chelation is known to decrease integrin expression and is often used as a positive control for inhibition in experiments [112,134]. Divalent cation stimulation of neutrophils with manganese or higher concentrations of magnesium induces the high-affinity conformation of β2-integrins without activating the neutrophil itself or increasing β2-integrin surface expression [140]. Through the manipulation of cation concentrations, Spillmann et al. demonstrated that β2-integrins must be in their active/high-affinity states to mediate adhesion [140]. A complete understanding of how divalent cations impact neutrophil function and integrin activation is also useful for interpreting clinical information following certain treatments. For example, magnesium sulfate treatment for preterm birth impairs neonatal innate immune cell recruitment and β2-integrin-dependent neutrophil responses [141].

3.4. Common Ligands

3.4.1. Recombinant ICAM-1

β2-integrins bind to intercellular adhesion molecules (e.g., ICAM-1) expressed on the surface of endothelial cells to transmigrate from the vasculature into inflamed tissues. This binding interaction induces outside-in activation and signaling of neutrophil β2-integrins. One of the most utilized ligands for understanding β2-integrin activation and signaling is ICAM-1 because it is a powerful tool for modeling physiologically relevant neutrophil interactions. In vitro, ICAM-1 stimulates neutrophil activation and adhesion in shear flow assays and even induces the clustering of neutrophil β2-integrins [33,45,108,112]. Although the usage of recombinant ICAM-1 is extremely common, there have been several discrepancies in the literature surrounding the nomenclature and usage of this ligand. Recombinant ICAM-1/Fc is often used interchangeably with recombinant ICAM-1. Based on our own observations (unpublished findings) and those cited in the literature, the Fc domain of ICAM-1/Fc is likely engaging Fc rectors on neutrophils and causing an inside-out activation cascade [122,133]. In light of this finding, we suggest that the choice of ICAM-1 construct is critical for the appropriate design of experiments interrogating neutrophil inside-out or outside-in activation, or both.

3.4.2. Fibrinogen

Fibrinogen is a glycoprotein found in the blood that is enzymatically converted to fibrin to promote clotting after damage occurs to vasculature or tissues. Neutrophil β2-integrins bind fibrinogen at sites of inflammation; therefore, it is used in vitro to determine integrin-dependent responses [33,113,142,143,144,145]. Lowell et al. determined that Src family kinases were important for β2 and β3-integrin signaling by evaluating hck−/− fgr−/− double mutant murine neutrophils on fibrinogen [142]. The double mutant neutrophils failed to spread on fibrinogen, but PMA stimulation was able to overcome the defect, indicating that Src kinases act upstream of PKC during integrin-mediated signaling. Fibrinogen initiates outside-in integrin signaling in both β2 and β3-integrins [145,146]; therefore, experiments utilizing this ligand for β2-integrins must rule out effects caused by β3-integrin engagement as well.

3.4.3. PolyRGD

β-integrins bind extracellular matrix proteins (e.g., fibrinogen, fibronectin, collagen, and von Willebrand factor) via their RGD (arginine-glycine-aspartic acid—RGD) site [147]. PolyRGD is a synthetic tripeptide used to engage integrins, and it is known for producing a robust CD18-dependent respiratory burst [54,148]. Other investigations found that Fc receptor knockout mice had decreased respiratory burst in response to polyRGD [149], suggesting that inside-out activation via Fc receptors, or Fc receptor cooperation, may also play a role in neutrophil responses to polyRGD. Because RGD binding sites exist on all β-integrins, polyRGD stimulates β1, β2, and β3-integrins expressed on neutrophils [84,150,151]. Therefore, the PolyRGD may not the best tool for isolating β2-integrin activation and signaling. However, it could be a useful tool for investigators interested in redundancy or cross-talk between parallel β-integrin activation and cell signaling cascades.

3.4.4. iC3b

Complement C3 fragment iC3b is a component of the complement system formed when complement factor I cleaves C3b. β2-integrins bind iC3b and are recognized as complement receptor 3 (CR3) [152]. In a physiological context, iC3b is an opsonin to support β2-integrin-mediated phagocytosis of pathogens [25]. Xue et al. used iC3b to stimulate K562 kindlin-3 knockdown cells and determined that kindlin-3 is required for iC3b-mediated outside-in β2-integrin signaling [107]. IC3b can also be used as a coating for neutrophil adhesion or in shear flow experiments [107]. Assays using iC3b as a tool are likely most relevant for in vitro modeling of diseases that may have iC3b-containing immune complex deposition contributing to neutrophil aggregate formation, such as Systemic Lupus Erythematosus (SLE) [12].

3.5. Assays

3.5.1. Flow Cytometry

Flow cytometry is a high-throughput technology that analyzes single cells from bulk populations. The technology detects and measures physical and chemical characteristics based on cell size and fluorescence. Neutrophils can be easily distinguished using flow cytometry based on their size and high granularity determined by a high side scatter profile when evaluating both side and forward scatter measurements. The usefulness of flow cytometry is widely known across many fields of research, and leukocyte researchers have also harnessed this powerful tool to assess β2-integrins. Using fluorescently labeled antibodies, researchers can measure the expression, avidity, and affinity of β2-integrins to learn more about how a protein or inhibitor impacts expression or to determine whether certain diseases cause changes in β2-integrin expression. Flow cytometry can also be used to measure neutrophil binding to ligands, such as ICAM-1, in the presence of pharmacological inhibitors or when isolated from transgenic mice [153,154]. Flow cytometry analysis of integrin expression is a relatively easy but powerful assay to complement other experiments. One caution is that neutrophil populations may exhibit autofluorescence, including autofluorescence attributed to contaminating eosinophils [85,155,156]. Non-specific staining can also occur if excessive concentrations of labeled antibodies are used, illustrating the importance of concentration optimization [157]. A significant benefit of flow cytometry is that multiple aspects of the neutrophil can be evaluated at once, such as integrin expression and cell viability. Imaging flow cytometry is a newer methodology used in neutrophil research. Specifically, this technology can be used to measure the fluorescence and morphology of neutrophils during functions, such as phagocytosis [158]. Because of its higher power in cellular analyses, this technique provides a breadth of information about cells. However, large data files can create challenges for data management and analysis [159].

3.5.2. Static Adhesion

Static adhesion is a common assay that has been used to assess neutrophils for over twenty years. Fluorescently labeled (e.g., calcein AM) neutrophils are added to ligand-covered plates for a designated time followed by subsequent washing and fluorescence readings [123,160]. This assay can evaluate the adhesion of neutrophil and neutrophil-like cells on most ligands, including human umbilical vein endothelial cell (HUVEC) monolayers [161,162]. Unfortunately, static adhesion assays are subject to technical variability due to the inversion procedure to “dump” cells. Further, static adhesion assays cannot fully recapitulate the physiological environment that occurs during shear flow adhesion. For example, neutrophil migration and adhesion under static adhesion require vinculin; however, vinculin was not required for integrin-mediated migration and adhesion when neutrophils were examined under shear flow [115]. Despite these differences, static adhesion assays do offer a high-throughput means to examine β2 integrin-mediated firm adhesion [66,123].

3.5.3. FRET

Förster Resonance Energy Transfer (FRET) (also referred to as Fluorescence Resonance Energy Transfer) is a method that shows energy transfer between two light-sensitive molecules based on distance [163]. With this newer technology, two proteins of interest can be labeled to quantitatively measure the interactions between the proteins. FRET can detect neutrophil β2-integrin conformational changes in the extracellular domain and the cytoplasmic tail when the α and β chains are labeled separately. Lefort et al. used this method to determine how inside-out activation of Mac-1 results in integrin headpiece extension from the bent conformation. Cytoplasmic domain FRET in K562 cells demonstrated that Mac-1 binding to ICAM-1 resulted in the separation of integrin cytoplasmic tails [108]. The sensitivity of this method has made it easier to determine protein interactions within living cells, including how integrins respond during neutrophil stimulation.

3.5.4. Integrin Crosslinking

Integrin crosslinking is a technique where anti-integrin antibodies (e.g., anti-CD18 mAb) are coated on a plate and used as the stimulus for β2-integrin activation and signaling [148]. This approach has historically been used to induce outside-in signaling of integrins. However, Jakus and colleagues determined that there was cooperative interaction between FcγRIIa and integrins during integrin crosslinking with anti-CD18 mAb due to the presence of Fc domains on mAbs. In this study, they demonstrated that anti-CD18 mAb crosslinking resulted in neutrophil respiratory burst. When anti-CD18 F(ab’)2 was used instead, respiratory burst no longer occurred despite significant neutrophil adhesion. This finding helped to prove that full neutrophil activation resulting in respiratory burst requires more than just integrin-ligand binding and outside-in β2-integrin activation [122]. This technique provides many benefits to understanding the cooperative signaling of integrins and FcγRs, but researchers should elect to use F(ab) or F(ab’)2 fragments, rather than intact antibodies, when trying to limit their stimulation to outside-in β2-integrin activation.

3.5.5. Flow Chamber Assays

Neutrophils are migratory cells where dynamic motion is a critical part of their function. Many in vitro neutrophil assays are unable to capture the dynamic process of neutrophil diapedesis. Flow chamber experiments can determine neutrophil crawling velocity, arrest, polarization, migration patterns, and diapedesis using microscopy. This technique offers a multitude of options for the use of ligands, cell type (whole blood, primary or differentiated neutrophil-like), chemoattractants, function-blocking antibodies, and immunofluorescence microscopy [54,137,164]. Flow chambers can also be coated with desired ligands and perfused with whole blood via tubing directly attached to murine carotid arteries. With this approach, Zarbock and colleagues demonstrated that E-selectin engagement is required for LFA-1-dependent rolling on ICAM-1 [165]. Microfluidic systems are also often used to determine the strength of neutrophil adhesion in relation to the ligand or a known amount of tension [137,166]. Morikis et al. demonstrated that neutrophils had increased adhesion and calcium flux in response to higher tension ligands while under shear flow [166]. Their study highlighted how high-affinity neutrophil β2-integrins recognize different levels of shear stress and tension and modulate downstream function and signaling to correspond to the stimulus [166].
In addition to the ligands utilized in shear flow assays, neutrophil interactions with cell monolayers can also be evaluated. Sule and colleagues demonstrated that neutrophils from patients with antiphospholipid syndrome display increased adhesion to HUVECs due to upregulated β2-integrin activation [126]. This system can also be designed to model organ-specific neutrophil interactions, such as blood-brain barrier inflammation. Gorina et al. showed that neutrophils use β2-integrins to crawl on ICAM-1 prior to diapedesis across isolated primary mouse brain microvascular endothelial cells [23]. In summary, the advantage of shear flow assays is that they offer a multitude of options to model healthy and diseased states.

3.5.6. Immunoblotting and Co-Immunoprecipitation

As the field continues to push toward effective drugs for targeting β2-integrins, we must consider other proteins that may serve as therapeutic targets. Many of the assays already discussed can be adapted using target-specific inhibitors to probe the involvement of individual proteins in β2-integrin activation and function. Another useful method is immunoblotting, which continues to be a frequently utilized approach for determining specific cell signaling patterns. A significant portion of our understanding of integrin signaling comes from immunoblotting experiments. Through immunoblotting, key signaling molecules downstream of integrin activation, such as Syk, have been identified [148]. Specifically, Lefort et al. demonstrated that Mac-1 outside-in activation with ICAM-1 activates only the Akt apoptosis regulatory pathway and not the p38 MAPK pathway [108]. With immunoblotting, researchers can determine the signaling mechanism that underlies a given function [21].
Co-immunoprecipitation (Co-IP) assays are used to identify the protein–protein interactions occurring within cells by indirectly capturing proteins that are bound to specific target proteins [167]. The unknown proteins are then evaluated using traditional immunoblotting techniques. Co-IP has been useful for determining binding partners of β2-integrins in both neutrophils and lymphocytes [45,168]. Co-IP can also be used to determine CD18 binding partners on the surface of neutrophils [130]. Through Co-IP experiments, Willeke et al. demonstrated that Syk binds to CD18 in fibrinogen-stimulated neutrophils, expanding the understanding of Syk’s role in β2-integrin activation and signaling [66].

3.5.7. Microscopy

Since the mid-1900s, researchers have utilized microscopy to evaluate neutrophils. The various methods of microscopy have aided researchers in their understanding of neutrophils, specifically how neutrophils change shape and polarize upon stimulation [164,169]. Confocal and immunofluorescence microscopy are the more commonly used microscopy platforms for evaluating neutrophils. Both methods utilize fluorescently labeled antibodies to evaluate β2-integrin clustering and surface distribution, subcellular localization, and colocalization with other proteins, such as F-actin [170,171,172,173].
Electron microscopy can be used to determine the binding of ligands to specific integrins. For example, Xu et al. used negative-stain electron microscopy to determine how the integrins αMβ2 and αXβ2 bind to iC3b, demonstrating that the different integrins bind to unique sites on iC3b [174]. Negative-stain electron microscopy can also be used to determine conformational changes of β2-integrins, deciphering between extended closed and extended open integrin conformations [135].
In a recent study, Wen et al. used high-resolution quantitative dynamic footprinting (qDF) microscopy, which is a total internal reflective fluorescence (TIRF)-based method, to analyze the relationship between kindlin-3 and β2-integrin activation in a shear flow assay of differentiated neutrophil-like HL60s [36]. Utilization of this method also demonstrated that integrins can obtain a high-affinity conformation (H+) without becoming extended (E) while rolling along ICAM-1 [175]. This EH+ conformation results in decreased neutrophil adhesion under flow. These findings were particularly interesting because they indicated an endogenous anti-inflammatory mechanism that could be harnessed by new integrin-targeting therapies [175].

3.6. In Vivo Experiments

While a vast amount of neutrophil β2-integrin research has been conducted in vitro, the use of mouse models has also made a significant impact on the field. The in vitro experiments utilizing primary cells or cell lines are essential for building a fundamental understanding of β2-integrins; however, in vitro findings are not always consistent with in vivo results. For example, in vitro experiments have consistently identified β2-integrins as essential receptors for neutrophil migration and adhesion, while more recent in vivo experiments have shown that the need for β2-integrins in vivo is variable. Table 5 summarizes a selection of in vivo models where β2-integrin dependence may vary depending on the stimulus or the organ in question. These differences are also noted when comparing leukocyte migration within a 3D collagen matrix versus placed on top of a collagen matrix [176,177]. The current understanding of these differences is that neutrophils are flexible in their responses to their environment. Migration on a 2D surface depends on cellular adhesion while movement within a 3D network, like collagen, depends on actomyosin contraction or actin polymerization [176,178,179,180]. What this does not explain is why pneumonia-causing pathogens have differential dependence on β2-integrins for neutrophil migration (Table 5) [181] or why certain β2-integrins are required while others are not [182]. These studies likely point toward evidence that the activation of internal cellular pathways following DAMP/PAMP recognition is also variable [183,184]. The recent development of humanized β2-integrin knockin mice should allow for better evaluation of integrin requirements in these disease models due to its ability to evaluate β2-integrin activation states in vivo [139].

Intravital Microscopy

Intravital microscopy in the mouse cremaster muscle is a well-established method for examining neutrophil function, characteristics, and interactions in the blood vessels. Leukocyte recruitment can be visualized in a variety of scenarios, including chemokine stimulation, fluorescently labeled leukocytes or transgenic mice expressing a fluorescent protein, and/or the application of pharmacological inhibitors [165,193,194]. Phillipson and colleagues used intravital microscopy to demonstrate the functional differences in LFA-1 and Mac-1 during MIP-2-induced neutrophil recruitment. Specifically, they found that LFA-1 is responsible for neutrophil adhesion while Mac-1 was responsible for neutrophil crawling [194]. This model was also used in experiments showing that E-selectin-induced slow rolling of neutrophils was LFA-1 dependent and Mac-1 independent [165]. As technology advances, researchers have expanded the field of intravital microscopy. Park and colleagues developed an intravital lung imaging system to examine neutrophil recruitment during sepsis-induced acute lung injury (ALI). In this study, investigators determined that decreased pulmonary microcirculation is due to obstructions of clustered neutrophils. Further, these neutrophils had high levels of surface Mac-1 expression, and the application of a Mac-1 inhibitor decreased sequestration in the pulmonary microvasculature [44]. Lim and colleagues also developed an advantageous model for evaluating β2-integrins. They generated a knockin mouse strain expressing CD11b conjugated to a monomeric yellow fluorescent protein (mYFP). This model can be utilized to image CD11b expressing cells in live mice or evaluate cell populations for CD11b expression in vitro. Although this approach results in all cells expressing CD11b to be YFP positive, this model allows for the analysis of functionally competent neutrophils in vivo without compromising β2-integrin–ligand interactions due to antibody binding [195]. Because of the breadth of options for examining neutrophils using intravital microscopy, this technique is an excellent next step for researchers wanting to translate in vitro findings into organ-specific in vivo scenarios [196].

4. Perspectives on the Study of Neutrophil β-Integrins

FcγRs often cooperate with β2-integrins in a complex mechanism of neutrophil activation and function. β2-integrin stimulation often results in neutrophil FcγR activation as well. For example, polyRGD is used to stimulate β2-integrins; however, neutrophils isolated from FcγR−/− mice have diminished respiratory burst response to polyRGD stimulation, suggesting FcγR-cooperation with polyRGD activation and signaling [149]. Further, wild-type murine neutrophils stimulated with polyRGD displayed p38 MAPK phosphorylation, which is not activated when neutrophil β2-integrins, are stimulated in an outside-in manner using ICAM-1 [108,149]. Previous work by Jakus et al. demonstrated that stimulation of neutrophils with anti-integrin monoclonal antibodies requires both β2-integrins and FcγRs [122]. These findings are strengthened by the fact that many FcγR-stimulated neutrophil events are β2-integrin-dependent. For example, equine neutrophil adhesion and respiratory burst stimulated by low-density insoluble immune complexes are dependent on both FcγR and β2-integrins [123,124]. Other studies have also demonstrated the interconnectedness of FcγRs and β2-integrins [120,197]. These findings add a layer of difficulty to experiments designed for deciphering the intracellular signaling events exclusive to either FcγRs or β2-integrins.
Although this review has focused primarily on β2-integrins, neutrophils also express β1- and β3-integrins [7,198]. The majority of β-integrin research in neutrophils has focused on β2-integrins; however, these less-studied integrins may also play significant roles in neutrophil activation and signaling [14]. Lomakina and Waugh demonstrated that neutrophils significantly adhere to vascular cell adhesion molecule 1 (VCAM-1) through the integrin α4β1 [199]. Further, there is evidence to demonstrate that engagement of β3-integrins and β1-integrins activate β2-integrins [198,200,201]. While the presence of these “other” β-integrins adds additional complexity to investigations focused on β2-integrin-exclusive signaling, they also represent an opportunity for additional research that may lead to a more complete, and even potentially clinically relevant, understanding of neutrophil β-integrin receptor functions in vivo.

5. Conclusions

β2-integrins have been the focus of intense research for decades, and many tools and assays have been developed to assess these receptors in neutrophils and neutrophil-like cells. These tools have been employed by researchers in very diverse fields, looking to decipher important questions regarding the biological function of β2-integrins. The combination of these cell types, tools, and assays offers a powerful resource to further understand β2-integrins and inform the future development of successful neutrophil targeting therapies.

Author Contributions

Conceptualization, H.E.C. and M.K.S.; writing—original draft preparation, H.E.C.; writing—review and editing, H.E.C. and M.K.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the following: USDA National Institute of Food and Agriculture (NIFA) Award 2018-67017-27632; NCSU College of Veterinary Medicine Department of Clinical Science Dissemination Fund; NCSU College of Veterinary Medicine Startup Funds (Sheats).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AKTserine/threonine kinase
CD18β chain of β2 integrins
CD11α chain of β2 integrins
CR3complement receptor 3
DAMPdamage-associated molecular pattern
dHL60differentiated HL60 cells
DMSOdimethylsulfoxide
eGFPenhanced green fluorescent protein
EVextracellular vesicle
FcγRFcgamma receptor
fMLPN-formylmethionine-leucyl-phenylalanine
GM-CSFgranulocyte-macrophage-colony-stimulating factor
GPCRG-protein-coupled receptor
iC3bcomplement protein fragment produced when complement factor I cleaves C3b
ICAM-1intercellular adhesion molecule 1
IL-8Rinterleukin 8 receptor
ITAMimmunoreceptor tyrosine-based activation motif
LDNlow-density neutrophil
LFA-1lymphocyte-function-associated antigen-1, αLβ2, CD11a/CD18
LPSlipopolysaccharide
LTB4leukotriene B4
Mac-1macrophage-1 antigen, αMβ2, CD11b/CD18
MIP-2macrophage inflammatory protein-2
NBTnitroblue tetrazolium
NDNnormal-density neutrophil
NETneutrophil extracellular trap
nEVneutrophil extracellular vesicle
PAMPpathogen-associated molecular pattern
PMAphorbol 12-myristate 13-acetate
PMNpolymorphonuclear leukocyte, neutrophil
PKCprotein kinase C
ROSreactive oxygen species
SykSpleen tyrosine kinase

References

  1. Abram, C.L.; Lowell, C.A. The Ins and Outs of Leukocyte Integrin Signaling. Annu. Rev. Immunol. 2009, 27, 339–362. [Google Scholar] [CrossRef] [PubMed]
  2. Bouti, P.; Webbers, S.D.S.; Fagerholm, S.C.; Alon, R.; Moser, M.; Matlung, H.L.; Kuijpers, T.W. β2 Integrin Signaling Cascade in Neutrophils: More Than a Single Function. Front. Immunol. 2021, 11, 619925. [Google Scholar] [CrossRef] [PubMed]
  3. Schymeinsky, J.; Mocsai, A.; Walzog, B. Neutrophil activation via β2 integrins (CD11/CD18): Molecular mechanisms and clinical implications. Thromb. Haemost. 2007, 98, 262–273. [Google Scholar] [CrossRef]
  4. Blythe, E.N.; Weaver, L.C.; Brown, A.; Dekaban, G.A. β2 Integrin CD11d/CD18: From Expression to an Emerging Role in Staged Leukocyte Migration. Front. Immunol. 2021, 12, 775447. [Google Scholar] [CrossRef] [PubMed]
  5. Montresor, A.; Toffali, L.; Constantin, G.; Laudanna, C.; Ley, K. Chemokines and the Signaling Modules Regulating Integrin Affinity. Front. Immunol. 2012, 3, 127. [Google Scholar] [CrossRef]
  6. Grabbe, S.; Wen, L.; Lyu, Q.; Ley, K.; Goult, B.T. Structural Basis of β2 Integrin Inside—Out Activation. Cells 2022, 11, 3039. [Google Scholar] [CrossRef]
  7. Williams, M.A.; Solomkin, J.S. Integrin-mediated signaling in human neutrophil functioning. J. Leukoc. Biol. 1999, 65, 725–736. [Google Scholar] [CrossRef] [PubMed]
  8. Jakus, Z.N.; Fodor, S.; Abram, C.L.; Lowell, C.A.; Mócsai, A. Immunoreceptor-like signaling by β2 and β3 integrins. Trends Cell Biol. 2007, 17, 493–501. [Google Scholar] [CrossRef]
  9. Schmidt, S.; Moser, M.; Sperandio, M. The molecular basis of leukocyte recruitment and its deficiencies. Mol. Immunol. 2013, 55, 49–58. [Google Scholar] [CrossRef]
  10. Yang, S.-C.; Tsai, Y.-F.; Pan, Y.-L.; Hwang, T.-L. Understanding the role of neutrophils in acute respiratory distress syndrome. Biomed. J. 2021, 44, 439–446. [Google Scholar] [CrossRef]
  11. Borges, L.; Pithon-Curi, T.C.; Curi, R.; Hatanaka, E. COVID-19 and Neutrophils: The Relationship between Hyperinflammation and Neutrophil Extracellular Traps. Mediat. Inflamm. 2020, 2020, 8829674. [Google Scholar] [CrossRef] [PubMed]
  12. Rosetti, F.; Mayadas, T.N. The many faces of Mac-1 in autoimmune disease. Immunol. Rev. 2016, 269, 175–193. [Google Scholar] [CrossRef] [PubMed]
  13. Rosen, H.; Law, S.K.A. The Leukocyte Cell Surface Receptor(s) for the iC3b Product of Complement. Curr. Top. Microbiol. Immunol. 1990, 153, 99–122. [Google Scholar] [CrossRef] [PubMed]
  14. Johnson, C.M.; O’brien, X.M.; Byrd, A.S.; Parisi, V.E.; Loosely, A.J.; Li, W.; Witt, H.; Faridi, H.M.; Lefort, C.T.; Gupta, V.; et al. Integrin Cross-Talk Regulates the Human Neutrophil Response to Fungal β-Glucan in the Context of the Extracellular Matrix: A Prominent Role for VLA3 in the Antifungal Response. J. Immunol. 2017, 198, 318–334. [Google Scholar] [CrossRef]
  15. Wright, S.D.; Levin, S.M.; Jong, M.T.C.; Chad, Z.; Kàbbashi, L.G. CR3 (CD11b/CD18) expresses one binding site for Arg-Gly-Asp-containing peptides and a second site for bacterial lipopolysaccharide. J. Exp. Med. 1989, 169, 175–183. [Google Scholar] [CrossRef]
  16. Mitroulis, I.; Alexaki, V.I.; Kourtzelis, I.; Ziogas, A.; Hajishengallis, G.; Chavakis, T. Leukocyte integrins: Role in leukocyte recruitment and as therapeutic targets in inflammatory disease. Pharmacol. Ther. 2015, 147, 123–135. [Google Scholar] [CrossRef]
  17. Zimmerman, T.; Blanco, F. Inhibitors Targeting the LFA-1/ICAM-1 Cell-Adhesion Interaction: Design and Mechanism of Action. Curr. Pharm. Des. 2008, 14, 2128–2139. [Google Scholar] [CrossRef]
  18. Raab-Westphal, S.; Marshall, J.F.; Goodman, S.L. Integrins as Therapeutic Targets: Successes and Cancers. Cancers 2017, 9, 110. [Google Scholar] [CrossRef]
  19. Matsumoto, K.; Kurasawa, T.; Yoshimoto, K.; Suzuki, K.; Takeuchi, T. Identification of neutrophil β2-integrin LFA-1 as a potential mechanistic biomarker in ANCA-associated vasculitis via microarray and validation analyses. Arthritis Res. Ther. 2021, 23, 136. [Google Scholar] [CrossRef]
  20. Teschner, D.; Cholaszczyńska, A.; Ries, F.; Beckert, H.; Theobald, M.; Grabbe, S.; Radsak, M.; Bros, M. CD11b Regulates Fungal Outgrowth but Not Neutrophil Recruitment in a Mouse Model of Invasive Pulmonary Aspergillosis. Front. Immunol. 2019, 10, 123. [Google Scholar] [CrossRef]
  21. Silva, J.C.; Rodrigues, N.C.; Thompson-Souza, G.A.; Muniz, V.D.S.; Neves, J.S.; Figueiredo, R.T. Mac-1 triggers neutrophil DNA extracellular trap formation to Aspergillus fumigatus independently of PAD4 histone citrullination. J. Leukoc. Biol. 2020, 107, 69–83. [Google Scholar] [CrossRef] [PubMed]
  22. Friedrichs, K.; Adam, M.; Remane, L.; Mollenhauer, M.; Rudolph, V.; Rudolph, T.K.; Andrié, R.P.; Stöckigt, F.; Schrickel, J.W.; Ravekes, T.; et al. Induction of Atrial Fibrillation by Neutrophils Critically Depends on CD11b/CD18 Integrins. PLoS ONE 2014, 9, e89307. [Google Scholar] [CrossRef] [PubMed]
  23. Gorina, R.; Lyck, R.; Vestweber, D.; Engelhardt, B. β2 Integrin–Mediated Crawling on Endothelial ICAM-1 and ICAM-2 Is a Prerequisite for Transcellular Neutrophil Diapedesis across the Inflamed Blood–Brain Barrier. J. Immunol. 2014, 192, 324–337. [Google Scholar] [CrossRef] [PubMed]
  24. Marchetti, L.; Engelhardt, B. Immune cell trafficking across the blood-brain barrier in the absence and presence of neuroinflammation. Vasc. Biol. 2020, 2, H1–H18. [Google Scholar] [CrossRef]
  25. Li, X.; Utomo, A.; Cullere, X.; Choi, M.M.; Milner, D.A., Jr.; Venkatesh, D.; Yun, S.-H.; Mayadas, T.N. The β-Glucan Receptor Dectin-1 Activates the Integrin Mac-1 in Neutrophils via Vav Protein Signaling to Promote Candida albicans Clearance. Cell Host Microbe 2011, 10, 603–615. [Google Scholar] [CrossRef]
  26. Woolhouse, I.S.; Bayley, D.L.; Lalor, P.; Adams, D.H.; Stockley, R.A. Endothelial interactions of neutrophils under flow in chronic obstructive pulmonary disease. Eur. Respir. J. 2005, 25, 612–617. [Google Scholar] [CrossRef]
  27. Blidberg, K.; Palmberg, L.; James, A.; Billing, B.; Henriksson, E.; Lantz, A.-S.; Larsson, K.; Dahlén, B. Adhesion molecules in subjects with COPD and healthy non-smokers: A cross sectional parallel group study. Respir. Res. 2013, 14, 47. [Google Scholar] [CrossRef]
  28. Overbeek, S.A.; Braber, S.; Henricks, P.A.J.; Kleinjan, M.; Kamp, V.M.; Georgiou, N.A.; Garssen, J.; Kraneveld, A.D.; Folkerts, G. Cigarette smoke induces β2-integrin-dependent neutrophil migration across human endothelium. Respir. Res. 2011, 12, 75. [Google Scholar] [CrossRef]
  29. Zou, J.; Chen, J.; Yan, Q.; Guo, Q.; Bao, C. Serum IL8 and mRNA level of CD11b in circulating neutrophils are increased in clinically amyopathic dermatomyositis with active interstitial lung disease. Clin. Rheumatol. 2016, 35, 117–125. [Google Scholar] [CrossRef]
  30. Edwards, D.N.; Bix, G.J. The Inflammatory Response After Ischemic Stroke: Targeting β2 and β1 Integrins. Front. Neurosci. 2019, 13, 540. [Google Scholar] [CrossRef]
  31. Yago, T.; Petrich, B.G.; Zhang, N.; Liu, Z.; Shao, B.; Ginsberg, M.H.; McEver, R.P. Blocking neutrophil integrin activation prevents ischemia–reperfusion injury. J. Exp. Med. 2015, 212, 1267–1281. [Google Scholar] [CrossRef] [PubMed]
  32. Dehnadi, A.; Cosimi, A.B.; Smith, R.N.; Li, X.; Alonso, J.L.; Means, T.K.; Arnaout, M.A. Prophylactic orthosteric inhibition of leukocyte integrin CD11b/CD18 prevents long-term fibrotic kidney failure in cynomolgus monkeys. Nat. Commun. 2017, 8, 13899. [Google Scholar] [CrossRef] [PubMed]
  33. Volmering, S.; Block, H.; Boras, M.; Lowell, C.A.; Zarbock, A. The Neutrophil Btk Signalosome Regulates Integrin Activation during Sterile Inflammation. Immunity 2016, 44, 73–87. [Google Scholar] [CrossRef] [PubMed]
  34. Chen, G.Y.; Nuñez, G. Sterile inflammation: Sensing and reacting to damage. Nat. Rev. Immunol. 2010, 10, 826–837. [Google Scholar] [CrossRef]
  35. Kuijpers, T.W.; Verhoeven, A.J.; Roos, D.; Kuijpers, T.W.; Van Lier, R.A.W.; Hamann, D.; De Boer, M.; Thung, L.Y.; Weening, R.S. Leukocyte adhesion deficiency type 1 (LAD-1)/variant. A novel immunodeficiency syndrome characterized by dysfunctional beta2 integrins. J. Clin. Investig. 1997, 100, 1725–1733. [Google Scholar] [CrossRef]
  36. Wen, L.; Marki, A.; Roy, P.; McArdle, S.; Sun, H.; Fan, Z.; Gingras, A.R.; Ginsberg, M.H.; Ley, K. Kindlin-3 recruitment to the plasma membrane precedes high-affinity β2-integrin and neutrophil arrest from rolling. Blood 2021, 137, 29–38. [Google Scholar] [CrossRef]
  37. Moser, M.; Bauer, M.; Schmid, S.; Ruppert, R.; Schmidt, S.; Sixt, M.; Wang, H.-V.; Sperandio, M.; Fässler, R. Kindlin-3 is required for β2 integrin–mediated leukocyte adhesion to endothelial cells. Nat. Med. 2009, 15, 300–305. [Google Scholar] [CrossRef]
  38. Daseke, M.J.; Chalise, U.; Becirovic-Agic, M.; Salomon, J.D.; Cook, L.M.; Case, A.J.; Lindsey, M.L. Neutrophil signaling during myocardial infarction wound repair. Cell. Signal 2021, 77, 109816. [Google Scholar] [CrossRef]
  39. Meisel, S.R.; Shapiro, H.; Radnay, J.; Neuman, Y.; Khaskia, A.-R.; Gruener, N.; Pauzner, H.; David, D. Increased Expression of Neutrophil and Monocyte Adhesion Molecules LFA-1 and Mac-1 and Their Ligand ICAM-1 and VLA-4 Throughout the Acute Phase of Myocardial Infarction: Possible Implications for Leukocyte Aggregation and Microvascular Plugging. J. Am. Coll. Cardiol. 1998, 31, 120–125. [Google Scholar] [CrossRef]
  40. Khawaja, A.A.; Pericleous, C.; Ripoll, V.M.; Porter, J.C.; Giles, I.P. Autoimmune rheumatic disease IgG has differential effects upon neutrophil integrin activation that is modulated by the endothelium. Sci. Rep. 2019, 9, 1283. [Google Scholar] [CrossRef]
  41. Simons, P.; Rinaldi, D.A.; Bondu, V.; Kell, A.M.; Bradfute, S.; Lidke, D.S.; Buranda, T. Integrin activation is an essential component of SARS-CoV-2 infection. Sci. Rep. 2021, 11, 1120398. [Google Scholar] [CrossRef] [PubMed]
  42. Narasaraju, T.; Tang, B.M.; Herrmann, M.; Muller, S.; Chow, V.T.K.; Radic, M. Neutrophilia and NETopathy as Key Pathologic Drivers of Progressive Lung Impairment in Patients With COVID-19. Front. Pharmacol. 2020, 11, 870. [Google Scholar] [CrossRef]
  43. Yuki, K.; Hou, L. Role of β2 Integrins in Neutrophils and Sepsis. Infect. Immun. 2020, 88, e00031-20. [Google Scholar] [CrossRef] [PubMed]
  44. Park, I.; Kim, M.; Choe, K.; Song, E.; Seo, H.; Hwang, Y.; Ahn, J.; Lee, S.-H.; Lee, J.H.; Jo, Y.H.; et al. Neutrophils disturb pulmonary microcirculation in sepsis-induced acute lung injury. Eur. Respir. J. 2019, 53, 1800786. [Google Scholar] [CrossRef]
  45. Xu, J.; Gao, X.-P.; Ramchandran, R.; Zhao, Y.-Y.; Vogel, S.M.; Malik, A.B. Nonmuscle myosin light-chain kinase mediates neutrophil transmigration in sepsis-induced lung inflammation by activating β2 integrins. Nat. Immunol. 2008, 9, 880–886. [Google Scholar] [CrossRef] [PubMed]
  46. Shimizu, K.; Libby, P.; Shubiki, R.; Sakuma, M.; Wang, Y.; Asano, K.; Mitchell, R.N.; Simon, D.I. Leukocyte Integrin Mac-1 Promotes Acute Cardiac Allograft Rejection. Circulation 2008, 117, 1997–2008. [Google Scholar] [CrossRef]
  47. Fagerholm, S.C.; MacPherson, M.; James, M.J.; Sevier-Guy, C.; Lau, C.S. The CD11b-integrin (ITGAM) and systemic lupus erythematosus. Lupus 2013, 22, 657–663. [Google Scholar] [CrossRef]
  48. Zhou, Y.; Wu, J.; Kucik, D.F.; White, N.B.; Redden, D.T.; Szalai, A.J.; Bullard, D.C.; Edberg, J.C. Multiple Lupus-AssociatedITGAMVariants Alter Mac-1 Functions on Neutrophils. Arthritis Rheum. 2013, 65, 2907–2916. [Google Scholar] [CrossRef] [PubMed]
  49. Wang, Y.; Gao, H.; Shi, C.; Erhardt, P.W.; Pavlovsky, A.; Soloviev, D.A.; Bledzka, K.; Ustinov, V.; Zhu, L.; Qin, J.; et al. Leukocyte integrin Mac-1 regulates thrombosis via interaction with platelet GPIbα. Nat. Commun. 2017, 8, 15559. [Google Scholar] [CrossRef] [PubMed]
  50. Berthold, T.; Glaubitz, M.; Muschter, S.; Groß, S.; Palankar, R.; Reil, A.; Helm, C.A.; Bakchoul, T.; Schwertz, H.; Bux, J.; et al. Human neutrophil antigen-3a antibodies induce neutrophil stiffening and conformational activation of CD11b without shedding of L-selectin. Transfusion 2015, 55, 2939–2948. [Google Scholar] [CrossRef]
  51. Henrich, D.; Zimmer, S.; Seebach, C.; Frank, J.; Barker, J.; Marzi, I. Trauma-Activated Polymorphonucleated Leukocytes Damage Endothelial Progenitor Cells: Probable role of CD11b/CD18-CD54 interaction and release of reactive oxygen specie. Shock 2011, 36, 216–222. [Google Scholar] [CrossRef] [PubMed]
  52. Hahm, E.; Li, J.; Kim, K.; Huh, S.; Rogelj, S.; Cho, J. Extracellular protein disulfide isomerase regulates ligand-binding activity of αMβ2 integrin and neutrophil recruitment during vascular inflammation. Blood 2013, 121, 3789–3800. [Google Scholar] [CrossRef]
  53. Goretti Riça, I.; Joughin, B.A.; Teke, M.E.; Emmons, T.R.; Griffith, A.M.B.; Cahill, L.A.; Banner-Goodspeed, V.M.M.; Robson, S.C.; Hernandez, J.M.; Segal, B.H.; et al. Neutrophil heterogeneity and emergence of a distinct population of CD11b/CD18-activated low-density neutrophils after trauma. J. Trauma Acute Care Surg. 2023, 94, 187–196. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, H.; Schaff, U.Y.; Green, C.E.; Chen, H.; Sarantos, M.R.; Hu, Y.; Wara, D.; Simon, S.I.; Lowell, C.A. Impaired Integrin-Dependent Function in Wiskott-Aldrich Syndrome Protein-Deficient Murine and Human Neutrophils. Immunity 2006, 25, 285–295. [Google Scholar] [CrossRef]
  55. Candotti, F. Clinical Manifestations and Pathophysiological Mechanisms of the Wiskott-Aldrich Syndrome. J. Clin. Immunol. 2018, 38, 13–27. [Google Scholar] [CrossRef]
  56. Pillay, J.; Braber, I.D.; Vrisekoop, N.; Kwast, L.M.; de Boer, R.J.; Borghans, J.A.M.; Tesselaar, K.; Koenderman, L. In vivo labeling with 2H2O reveals a human neutrophil lifespan of 5.4 days. Blood 2010, 116, 625–627. [Google Scholar] [CrossRef] [PubMed]
  57. Tak, T.; Tesselaar, K.; Pillay, J.; Borghans, J.A.M.; Koenderman, L. What’s your age again? Determination of human neutrophil half-lives revisited. J. Leukoc. Biol. 2013, 94, 595–601. [Google Scholar] [CrossRef] [PubMed]
  58. Wright, H.L.; Makki, F.A.; Moots, R.J.; Edwards, S.W. Low-density granulocytes: Functionally distinct, immature neutrophils in rheumatoid arthritis with altered properties and defective TNF signalling. J. Leukoc. Biol. 2017, 101, 599–611. [Google Scholar] [CrossRef]
  59. Quach, A.; Ferrante, A. The Application of Dextran Sedimentation as an Initial Step in Neutrophil Purification Promotes Their Stimulation, due to the Presence of Monocytes. J. Immunol. Res. 2017, 2017, 1254792. [Google Scholar] [CrossRef]
  60. Kuhns, D.B.; Priel, D.A.L.; Chu, J.; Zarember, K.A. Isolation and Functional Analysis of Human Neutrophils. Curr. Protoc. Immunol. 2016, 111, 7.23.1–7.23.16. [Google Scholar] [CrossRef]
  61. Jimbo, S.; Suleman, M.; Maina, T.; Prysliak, T.; Mulongo, M.; Perez-Casal, J. Effect of Mycoplasma bovis on bovine neutrophils. Veter Immunol. Immunopathol. 2017, 188, 27–33. [Google Scholar] [CrossRef] [PubMed]
  62. Villanueva, E.; Yalavarthi, S.; Berthier, C.C.; Hodgin, J.B.; Khandpur, R.; Lin, A.M.; Rubin, C.J.; Zhao, W.; Olsen, S.H.; Klinker, M.; et al. Netting Neutrophils Induce Endothelial Damage, Infiltrate Tissues, and Expose Immunostimulatory Molecules in Systemic Lupus Erythematosus. J. Immunol. 2012, 187, 538–552. [Google Scholar] [CrossRef] [PubMed]
  63. Schweizer, T.A.; Shambat, S.M.; Vulin, C.; Hoeller, S.; Acevedo, C.; Huemer, M.; Gomez-Mejia, A.; Chang, C.; Baum, J.; Hertegonne, S.; et al. Blunted sFasL signalling exacerbates TNF-driven neutrophil necroptosis in critically ill COVID-19 patients. Clin. Transl. Immunol. 2021, 10, e1357. [Google Scholar] [CrossRef] [PubMed]
  64. Cotter, M.J.; Norman, K.E.; Hellewell, P.G.; Ridger, V.C. A Novel Method for Isolation of Neutrophils from Murine Blood Using Negative Immunomagnetic Separation. Am. J. Pathol. 2001, 159, 473–481. [Google Scholar] [CrossRef] [PubMed]
  65. Son, K.; Mukherjee, M.; McIntyre, B.A.S.; Eguez, J.C.; Radford, K.; LaVigne, N.; Ethier, C.; Davoine, F.; Janssen, L.; Lacy, P.; et al. Improved recovery of functionally active eosinophils and neutrophils using novel immunomagnetic technology. J. Immunol. Methods 2017, 449, 44–55. [Google Scholar] [CrossRef]
  66. Willeke, T.; Schymeinsky, J.; Prange, P.; Zahler, S.; Walzog, B. A role for Syk-kinase in the control of the binding cycle of the β2 integrins (CD11/CD18) in human polymorphonuclear neutrophils. J. Leukoc. Biol. 2003, 74, 260–269. [Google Scholar] [CrossRef]
  67. Diacovo, T.G.; Roth, S.J.; Buccola, J.M.; Bainton, D.F.; Springer, T.A. Neutrophil Rolling, Arrest, and Transmigration Across Activated, Surface-Adherent Platelets Via Sequential Action of P-Selectin and the β2-Integrin CDllb/CDl8. Blood 1996, 88, 146–157. [Google Scholar] [CrossRef]
  68. Lowenthal, R.; Park, D.; Goldman, J.; Th’ng, K.; Hill, R.; Whyte, G. The cyropresevration of leukaemia cells: Morphological and functional changes. Br. J. Haematol. 1976, 34, 105–117. [Google Scholar] [CrossRef]
  69. Hill, R.S.; Still, B.J.; Mackinder, C.A. Improved functional recovery of human granulocytes after cryopreservation. Cryobiology 1981, 18, 533–540. [Google Scholar] [CrossRef]
  70. Malawista, S.E.; Van Blaricom, G.; Breitenstein, M.G. Cryopreservable neutrophil surrogates. Stored cytoplasts from human polymorphonuclear leukocytes retain chemotactic, phagocytic, and microbicidal function. J. Clin. Investig. 1989, 83, 728–732. [Google Scholar] [CrossRef]
  71. Voetman, A.; Bot, A.; Roos, D. Cryopreservation of enucleated human neutrophils (PMN cytoplasts). Blood 1984, 63, 234–237. [Google Scholar] [CrossRef] [PubMed]
  72. Hong, C.-W. Extracellular Vesicles of Neutrophils. Immune Netw. 2018, 18, e43. [Google Scholar] [CrossRef] [PubMed]
  73. Kolonics, F.; Szeifert, V.; Timár, C.I.; Ligeti, E.; Lőrincz, Á.M. The Functional Heterogeneity of Neutrophil-Derived Extracellular Vesicles Reflects the Status of the Parent Cell. Cells 2020, 9, 2718. [Google Scholar] [CrossRef] [PubMed]
  74. Zhou, Y.; Bréchard, S. Neutrophil Extracellular Vesicles: A Delicate Balance between Pro-Inflammatory Responses and Anti-Inflammatory Therapies. Cells 2022, 11, 3318. [Google Scholar] [CrossRef] [PubMed]
  75. Szeifert, V.; Kolonics, F.; Bartos, B.; Khamari, D.; Vági, P.; Barna, L.; Ligeti, E.; Lőrincz, M. Mac-1 Receptor Clustering Initiates Production of Pro-Inflammatory, Antibacterial Extracellular Vesicles from Neutrophils. Front. Immunol. 2021, 12, 671995. [Google Scholar] [CrossRef]
  76. Xu, Z.; Cai, J.; Gao, J.; White, G.C.; Chen, F.; Ma, Y.-Q. Interaction of kindlin-3 and β2-integrins differentially regulates neutrophil recruitment and NET release in mice. Blood 2015, 126, 373–377. [Google Scholar] [CrossRef]
  77. Wong, S.L.; Demers, M.; Martinod, K.; Gallant, M.; Wang, Y.; Goldfine, A.B.; Kahn, C.R.; Wagner, D.D. Diabetes primes neutrophils to undergo NETosis, which impairs wound healing. Nat. Med. 2015, 21, 815–819. [Google Scholar] [CrossRef]
  78. Alder, M.N.; Mallela, J.; Opoka, A.M.; Lahni, P.; Hildeman, D.A.; Wong, H.R. Olfactomedin 4 marks a subset of neutrophils in mice. Innate Immun. 2018, 25, 22–33. [Google Scholar] [CrossRef]
  79. Helou, D.G.; Braham, S.; De Chaisemartin, L.; Granger, V.; Damien, M.-H.; Pallardy, M.; Kerdine-Römer, S.; Chollet-Martin, S. Nrf2 downregulates zymosan-induced neutrophil activation and modulates migration. PLoS ONE 2019, 14, e0216465. [Google Scholar] [CrossRef]
  80. Rivadeneyra, L.; Charó, N.; Kviatcovsky, D.; de la Barrera, S.; Gómez, R.M.; Schattner, M. Role of neutrophils in CVB3 infection and viral myocarditis. J. Mol. Cell. Cardiol. 2018, 125, 149–161. [Google Scholar] [CrossRef]
  81. Mestas, J.; Hughes, C.C.W. Of mice and not men: Differences between mouse and human immunology. J. Immunol. 2004, 172, 2731–2738. [Google Scholar] [CrossRef]
  82. Bruhns, P.; Jönsson, F. Mouse and human FcR effector functions. Immunol. Rev. 2015, 268, 25–51. [Google Scholar] [CrossRef] [PubMed]
  83. Castro-Dopico, T.; Clatworthy, M.R. IgG and Fcγ Receptors in Intestinal Immunity and Inflammation. Front. Immunol. 2019, 10, 805. [Google Scholar] [CrossRef] [PubMed]
  84. Bertram, A.; Zhang, H.; von Vietinghoff, S.; de Pablo, C.; Haller, H.; Shushakova, N.; Ley, K. Protein Kinase C-θ Is Required for Murine Neutrophil Recruitment and Adhesion Strengthening under Flow. J. Immunol. 2012, 188, 4043–4051. [Google Scholar] [CrossRef] [PubMed]
  85. Dorward, D.A.; Lucas, C.D.; Alessandri, A.L.; Marwick, J.A.; Rossi, F.; Dransfield, I.; Haslett, C.; Dhaliwal, K.; Rossi, A.G. Technical Advance: Autofluorescence-based sorting: Rapid and nonperturbing isolation of ultrapure neutrophils to determine cytokine production. J. Leukoc. Biol. 2013, 94, 193–202. [Google Scholar] [CrossRef] [PubMed]
  86. Grieshaber-Bouyer, R.; Nigrovic, P.A. Neutrophil Heterogeneity as Therapeutic Opportunity in Immune-Mediated Disease. Front. Immunol. 2019, 10, 346. [Google Scholar] [CrossRef] [PubMed]
  87. Silvestre-Roig, C.; Fridlender, Z.G.; Glogauer, M.; Scapini, P. Neutrophil Diversity in Health and Disease. Trends Immunol. 2019, 40, 565–583. [Google Scholar] [CrossRef] [PubMed]
  88. Eruslanov, E.B.; Singhal, S.; Albelda, S.M. Mouse versus Human Neutrophils in Cancer: A Major Knowledge Gap. Trends Cancer 2017, 3, 149–160. [Google Scholar] [CrossRef]
  89. Soroush, F.; Tang, Y.; Mustafa, O.; Sun, S.; Yang, Q.; Kilpatrick, L.E.; Kiani, M.F. Neutrophil-endothelial interactions of murine cells is not a good predictor of their interactions in human cells. FASEB J. 2020, 34, 2691–2702. [Google Scholar] [CrossRef]
  90. Ecker, S.; Chen, L.; Pancaldi, V.; Bagger, F.O.; Fernández, J.M.; Carrillo de Santa Pau, E.C.; Juan, D.; Mann, A.L.; Watt, S.; Casale, F.P.; et al. Genome-wide analysis of differential transcriptional and epigenetic variability across human immune cell types. Genome Biol. 2017, 18, 18. [Google Scholar] [CrossRef]
  91. Atallah-Yunes, S.A.; Ready, A.; Newburger, P.E. Benign ethnic neutropenia. Blood Rev. 2019, 37, 100586. [Google Scholar] [CrossRef] [PubMed]
  92. Saul, M.C.; Philip, V.M.; Reinholdt, L.G.; Chesler, E.J. High-Diversity Mouse Populations for Complex Traits. Trends Genet. 2019, 35, 501–514. [Google Scholar] [CrossRef] [PubMed]
  93. Gupta, S.; Lee, C.-M.; Wang, J.-F.; Parodo, J.; Jia, S.-H.; Hu, J.; Marshall, J.C. Heat-shock protein-90 prolongs septic neutrophil survival by protecting c-Src kinase and caspase-8 from proteasomal degradation. J. Leukoc. Biol. 2018, 103, 933–944. [Google Scholar] [CrossRef] [PubMed]
  94. Liu, Y.G.; Teng, Y.S.; Cheng, P.; Kong, H.; Lv, P.Y.; Mao, F.Y.; Wu, X.L.; Hao, C.J.; Chen, W.; Yang, S.M.; et al. Abrogation of cathepsin C by Helicobacter pylori impairs neutrophil activation to promote gastric infection. FASEB J. 2019, 33, 5018–5033. [Google Scholar] [CrossRef] [PubMed]
  95. Hickstein, D.D.; Smith, A.; Fisher, W.; Beatty, P.G.; Schwartz, B.R.; Harlan, J.M.; Root, R.K.; Locksley, R.M. Expression of leukocyte adherence-related glycoproteins during differentiation of HL-60 promyelocytic leukemia cells. J. Immunol. 1987, 138, 513–519. [Google Scholar] [CrossRef]
  96. Gupta, D.; Shah, H.P.; Malu, K.; Berliner, N.; Gaines, P. Differentiation and Characterization of Myeloid Cells. Curr. Protoc. Immunol. 2014, 104, 22F.5.1–22F.5.28. [Google Scholar] [CrossRef]
  97. Blanter, M.; Gouwy, M.; Struyf, S. Studying Neutrophil Function in vitro: Cell Models and Environmental Factors. J. Inflamm. Res. 2021, 14, 141–162. [Google Scholar] [CrossRef]
  98. Deevi, R.K.; Koney-Dash, M.; Kissenpfennig, A.; Johnston, J.A.; Schuh, K.; Walter, U.; Dib, K. Vasodilator-Stimulated Phosphoprotein Regulates Inside-Out Signaling of β2 Integrins in Neutrophils. J. Immunol. 2010, 184, 6575–6584. [Google Scholar] [CrossRef]
  99. Rincón, E.; Rocha-Gregg, B.L.; Collins, S.R. A map of gene expression in neutrophil-like cell lines. BMC Genom. 2018, 19, 573. [Google Scholar] [CrossRef]
  100. Sham, R.L.; Phatak, P.D.; Belanger, K.A.; Packman, C.H. Functional properties of HL60 cells matured with all-trans-retinoic acid and DMSO: Differences in response to interleukin-8 and fMLP. Leuk. Res. 1995, 19, 1–6. [Google Scholar] [CrossRef]
  101. Launay, S.; Brown, G.; Machesky, L.M. Expression of WASP and Scar1/WAVE1 actin-associated proteins is differentially modulated during differentiation of HL-60 cells. Cell Motil. Cytoskelet. 2003, 54, 274–285. [Google Scholar] [CrossRef] [PubMed]
  102. Hauert, A.B.; Martinelli, S.; Marone, C.; Niggli, V. Differentiated HL-60 cells are a valid model system for the analysis of human neutrophil migration and chemotaxis. Int. J. Biochem. Cell Biol. 2002, 34, 838–854. [Google Scholar] [CrossRef] [PubMed]
  103. Patcha, V.; Wigren, J.; Winberg, M.E.; Rasmusson, B.; Li, J.; Särndahl, E. Differential inside-out activation of β2-integrins by leukotriene B4 and fMLP in human neutrophils. Exp. Cell Res. 2004, 300, 308–319. [Google Scholar] [CrossRef] [PubMed]
  104. Lubbert, M.; Herrmann, F.; Koeffler, H.P. Expression and regulation of myeloid-specific genes in normal and leukemic myeloid cells. Blood 1991, 77, 909–924. [Google Scholar] [CrossRef] [PubMed]
  105. Weber, C.; Lu, C.F.; Casasnovas, J.M.; Springer, T.A. Role of alpha L beta 2 integrin avidity in transendothelial chemotaxis of mononuclear cells. J. Immunol. Methods 1997, 159, 3968–3975. [Google Scholar] [CrossRef]
  106. DuMont, A.L.; Yoong, P.; Day, C.J.; Alonzo, F.; McDonald, W.H.; Jennings, M.P.; Torres, V.J. Staphylococcus aureus LukAB cytotoxin kills human neutrophils by targeting the CD11b subunit of the integrin Mac-1. Proc. Natl. Acad. Sci. USA 2013, 110, 10794–10799. [Google Scholar] [CrossRef] [PubMed]
  107. Xue, Z.-H.; Feng, C.; Liu, W.-L.; Tan, S.-M. A Role of Kindlin-3 in Integrin αMβ2 Outside-In Signaling and the Syk-Vav1-Rac1/Cdc42 Signaling Axis. PLoS ONE 2013, 8, e56911. [Google Scholar] [CrossRef]
  108. Lefort, C.T.; Hyun, Y.-M.; Schultz, J.B.; Law, F.-Y.; Waugh, R.E.; Knauf, P.A.; Kim, M. Outside-In Signal Transmission by Conformational Changes in Integrin Mac-1. J. Immunol. 2009, 183, 6460–6468. [Google Scholar] [CrossRef]
  109. Harlan, J.M.; Winn, R.K. Leukocyte-endothelial interactions: Clinical trials of anti-adhesion therapy. Crit. Care Med. 2002, 30, S214–S219. [Google Scholar] [CrossRef]
  110. Dove, A. CD18 trials disappoint again. Nat. Biotechnol. 2000, 18, 817–818. [Google Scholar] [CrossRef]
  111. Hafeez Faridi, M.; Altintas, M.M.; Gomez, C.; Camilo Doque, J.; Vazquez-Padron, R.I.; Gupta, V. Small molecular agonists of integrin CD11b/CD18 do not induce global conformational changes and are significantly better than activating antibodies in reducing vascular injury. Biochim. Biophys. Acta 2013, 1830, 3696–3710. [Google Scholar] [CrossRef] [PubMed]
  112. Celik, E.; Faridi, M.H.; Kumar, V.; Deep, S.; Moy, V.T.; Gupta, V. Agonist Leukadherin-1 Increases CD11b/CD18-Dependent Adhesion Via Membrane Tethers. Biophys. J. 2013, 105, 2517–2527. [Google Scholar] [CrossRef] [PubMed]
  113. Chu, J.Y.; McCormick, B.; Mazelyte, G.; Michael, M.; Vermeren, S. HoxB8 neutrophils replicate Fcγ receptor and integrin-induced neutrophil signaling and functions. J. Leukoc. Biol. 2019, 105, 93–100. [Google Scholar] [CrossRef] [PubMed]
  114. Saul, S.; Castelbou, C.; Fickentscher, C.; Demaurex, N. Signaling and functional competency of neutrophils derived from bone-marrow cells expressing the ER-HOXB8 oncoprotein. J. Leukoc. Biol. 2019, 106, 1101–1115. [Google Scholar] [CrossRef]
  115. Wilson, Z.S.; Witt, H.; Hazlett, L.; Harman, M.; Neumann, B.M.; Whitman, A.; Patel, M.; Ross, R.S.; Franck, C.; Reichner, J.S.; et al. Context-Dependent Role of Vinculin in Neutrophil Adhesion, Motility and Trafficking. Sci. Rep. 2020, 10, 2142. [Google Scholar] [CrossRef]
  116. Cohen, J.T.; Danise, M.; Hinman, K.D.; Neumann, B.M.; Johnson, R.; Wilson, Z.S.; Chorzalska, A.; Dubielecka, P.M.; Lefort, C.T. Engraftment, Fate, and Function of HoxB8-Conditional Neutrophil Progenitors in the Unconditioned Murine Host. Front. Cell Dev. Biol. 2022, 10, 840894. [Google Scholar] [CrossRef]
  117. Bromberger, T.; Klapproth, S.; Rohwedder, I.; Weber, J.; Pick, R.; Mittmann, L.; Min-Weißenhorn, S.J.; Reichel, C.A.; Scheiermann, C.; Sperandio, M.; et al. Binding of Rap1 and Riam to Talin1 Fine-Tune β2 Integrin Activity During Leukocyte Trafficking. Front. Immunol. 2021, 12, 702345. [Google Scholar] [CrossRef]
  118. Bromberger, T.; Klapproth, S.; Sperandio, M.; Moser, M. Humanized Beta2 Integrin-Expressing Hoxb8 Cells Serve as Model to Study Integrin Activation. Cells 2022, 11, 1532. [Google Scholar] [CrossRef]
  119. Penberthy, T.; Jiang, Y.; Luscinskas, F.; Graves, D. MCP-1-stimulated monocytes preferentially utilize β2-integrins to migrate on laminin and fibronectin. Am. J. Physiol. 1995, 269, C60–C68. [Google Scholar] [CrossRef]
  120. Graham, I.L.; Lefkowith, J.B.; Anderson, D.C.; Brown, E.J.; Leikowith, J.B.; Anderson, D.C.; Browntln, E.J. Immune complex-stimulated neutrophil LTB4 production is dependent on β2 integrins. J. Cell Biol. 1993, 120, 1509–1517. [Google Scholar] [CrossRef]
  121. Berton, G.; Fumagalli, L.; Laudanna, C.; Sorio, C. β2 Integrin-dependent Protein Tyrosine Phosphorylation and Activation of the FGR Protein Tyrosine Kinase in Human Neutrophils. J. Cell Biol. 1994, 126, 1111–1121. [Google Scholar] [CrossRef] [PubMed]
  122. Jakus, Z.; Berton, G.; Ligeti, E.; Lowell, C.A.; Mócsai, A. Responses of Neutrophils to Anti-Integrin Antibodies Depends on Costimulation through Low Affinity FcγRs: Full Activation Requires Both Integrin and Nonintegrin Signals. J. Immunol. 2004, 173, 2068–2077. [Google Scholar] [CrossRef] [PubMed]
  123. Sheats, M.K.; Pescosolido, K.C.; Hefner, E.M.; Sung, E.J.; Adler, K.B.; Jones, S.L. Myristoylated Alanine Rich C Kinase Substrate (MARCKS) is essential to β2-integrin dependent responses of equine neutrophils. Veter Immunol. Immunopathol. 2014, 160, 167–176. [Google Scholar] [CrossRef] [PubMed]
  124. Jones, S.L.; Sharief, Y.; Chilcoat, C.D. Signaling mechanism for equine neutrophil activation by immune complexes. Veter Immunol. Immunopathol. 2001, 82, 87–100. [Google Scholar] [CrossRef]
  125. Feng, C.; Zhang, L.; Almulki, L.; Faez, S.; Whitford, M.; Hafezi-Moghadam, A.; Cross, A.S. Endogenous PMN sialidase activity exposes activation epitope on CD11b/CD18 which enhances its binding interaction with ICAM-1. J. Leukoc. Biol. 2011, 90, 313–321. [Google Scholar] [CrossRef]
  126. Sule, G.; Kelley, W.J.; Gockman, K.; Yalavarthi, S.; Vreede, A.P.; Banka, A.L.; Bockenstedt, P.L.; Eniola-Adefeso, O.; Knight, J.S. Increased Adhesive Potential of Antiphospholipid Syndrome Neutrophils Mediated by β2 Integrin Mac-1. Arthritis Rheumatol. 2020, 72, 114–124. [Google Scholar] [CrossRef]
  127. Wilson, Z.S.; Ahn, L.B.; Serratelli, W.S.; Belley, M.D.; Lomas-Neira, J.; Sen, M.; Lefort, C.T. Activated β2 Integrins Restrict Neutrophil Recruitment during Murine Acute Pseudomonal Pneumonia. Am. J. Respir. Cell Mol. Biol. 2017, 56, 620–627. [Google Scholar] [CrossRef]
  128. Pollara, J.; Tay, M.Z.; Edwards, R.W.; Goodman, D.; Crowley, A.R.; Edwards, R.J.; Easterhoff, D.; Conley, H.E.; Hoxie, T.; Gurley, T.; et al. Functional Homology for Antibody-Dependent Phagocytosis Across Humans and Rhesus Macaques. Front. Immunol. 2021, 12, 678511. [Google Scholar] [CrossRef]
  129. Walzog, B.; Weinmann, P.; Jeblonski, F.; Scharffetter-Kochanek, K.; Bommert, K.; Gaehtgens, P. A role for β2 integrins (CD11/CD18) in the regulation of cytokine gene expression of polymorphonuclear neutrophils during the inflammatory response. FASEB J. 1999, 13, 1855–1865. [Google Scholar] [CrossRef]
  130. Burnett, A.; Gomez, I.; De Leon, D.D.; Ariaans, M.; Progias, P.; Kammerer, R.A.; Velasco, G.; Marron, M.; Hellewell, P.; Ridger, V. Angiopoietin-1 enhances neutrophil chemotaxis in vitro and migration in vivo through interaction with CD18 and release of CCL4. Sci. Rep. 2017, 7, 2332. [Google Scholar] [CrossRef]
  131. Kunkel, E.J.; Dunne, J.L.; Ley, K. Leukocyte Arrest During Cytokine-Dependent Inflammation In Vivo. J. Immunol. 2000, 164, 3301–3308. [Google Scholar] [CrossRef] [PubMed]
  132. Petruzzelli, L.; Maduzia, L.; Springer3, T.A. Activation of Lymphocyte Function-Associated Molecule4 (CDlla/CD18) and Mac-1 (CD11 b/CD18) Mimicked by an Antibody Directed Against CD18’. J. Immunol. 1995, 155, 854–866. [Google Scholar] [CrossRef] [PubMed]
  133. Jones, S.L.; Knaus, U.G.; Bokoch, G.M.; Brown, E.J. Two Signaling Mechanisms for Activation of αMβ2 Avidity in Polymorphonuclear Neutrophils. J. Biol. Chem. 1998, 273, 10556–10566. [Google Scholar] [CrossRef] [PubMed]
  134. Diamond, M.S.; Springer, T.A. A Subpopulation of Mac-1 (CDllb/CD18) Molecules Mediates Neutrophil Adhesion to ICAM-1 and Fibfinogen. J. Cell Biol. 1993, 120, 545–556. [Google Scholar] [CrossRef]
  135. Nishida, N.; Xie, C.; Shimaoka, M.; Cheng, Y.; Walz, T.; Springer, T.A. Activation of Leukocyte β2 Integrins by Conversion from Bent to Extended Conformations. Immunity 2006, 25, 583–594. [Google Scholar] [CrossRef]
  136. Fossati-Jimack, L.; Ling, G.S.; Cortini, A.; Szajna, M.; Malik, T.H.; McDonald, J.U.; Pickering, M.C.; Cook, H.T.; Taylor, P.R.; Botto, M. Phagocytosis Is the Main CR3-Mediated Function Affected by the Lupus-Associated Variant of CD11b in Human Myeloid Cells. PLoS ONE 2013, 8, e57082. [Google Scholar] [CrossRef]
  137. Sun, X.; Huang, B.; Pan, Y.; Fang, J.; Wang, H.; Ji, Y.; Ling, Y.; Guo, P.; Lin, J.; Li, Q.; et al. Spatiotemporal characteristics of P-selectin-induced β2 integrin activation of human neutrophils under flow. Front. Immunol. 2022, 13, 1023865. [Google Scholar] [CrossRef]
  138. Azcutia, V.; Kelm, M.; Luissint, A.-C.; Boerner, K.; Flemming, S.; Quiros, M.; Newton, G.; Nusrat, A.; Luscinskas, F.W.; Parkos, C.A. Neutrophil expressed CD47 regulates CD11b/CD18-dependent neutrophil transepithelial migration in the intestine in vivo. Mucosal Immunol. 2021, 14, 331–341. [Google Scholar] [CrossRef]
  139. Wen, L.; Marki, A.; Wang, Z.; Orecchioni, M.; Makings, J.; Billitti, M.; Wang, E.; Suthahar, S.S.A.; Kim, K.; Kiosses, W.B.; et al. A humanized β2 integrin knockin mouse reveals localized intra- and extravascular neutrophil integrin activation in vivo. Cell Rep. 2022, 39, 110876. [Google Scholar] [CrossRef]
  140. Spillmann, C.; Osorio, D.; Waugh, R. Integrin activation by divalent ions affects neutrophil homotypic adhesion. Ann. Biomed. Eng. 2002, 30, 1002–1011. [Google Scholar] [CrossRef]
  141. Mehta, R.; Petrova, A. Intrapartum Magnesium Sulfate Exposure Attenuates Neutrophil Function in Preterm Neonates. Neonatology 2006, 89, 99–103. [Google Scholar] [CrossRef] [PubMed]
  142. Lowell, C.A.; Fumagalli, L.; Berton, G. Deficiency of Src family kinases p59/61hck and p58c-fgr results in defective adhesion-dependent neutrophil functions. J. Cell Biol. 1996, 133, 895–910. [Google Scholar] [CrossRef] [PubMed]
  143. Zen, K.; Guo, Y.-L.; Li, L.-M.; Bian, Z.; Zhang, C.-Y.; Liu, Y. Cleavage of the CD11b extracellular domain by the leukocyte serprocidins is critical for neutrophil detachment during chemotaxis. Blood 2011, 117, 4885–4894. [Google Scholar] [CrossRef]
  144. Szczur, K.; Zheng, Y.; Filippi, M.-D. The small Rho GTPase Cdc42 regulates neutrophil polarity via CD11b integrin signaling. Blood 2009, 114, 4527–4537. [Google Scholar] [CrossRef] [PubMed]
  145. McMillan, S.J.; Sharma, R.S.; McKenzie, E.J.; Richards, H.E.; Zhang, J.; Prescott, A.; Crocker, P.R. Siglec-E is a negative regulator of acute pulmonary neutrophil inflammation and suppresses CD11b β2-integrin–dependent signaling. Blood 2013, 121, 2084–2094. [Google Scholar] [CrossRef]
  146. Cox, D.; Aoki, T.; Seki, J.; Motoyama, Y.; Yoshida, K. The pharmacology of the integrins. Med. Res. Rev. 1994, 14, 195–228. [Google Scholar] [CrossRef]
  147. Ruoslahti, E.; Pierschbacher, M.D. New Perspectives in Cell Adhesion: RGD and Integrins. Science 1987, 238, 491–497. [Google Scholar] [CrossRef]
  148. Mócsai, A.; Zhou, M.; Meng, F.; Tybulewicz, V.L.; Lowell, C.A. Syk Is Required for Integrin Signaling in Neutrophils. Immunity 2002, 16, 547–558. [Google Scholar] [CrossRef]
  149. Mócsai, A.; Abram, C.L.; Jakus, Z.; Hu, Y.; Lanier, L.L.; Lowell, C.A. Integrin signaling in neutrophils and macrophages uses adaptors containing immunoreceptor tyrosine-based activation motifs. Nat. Immunol. 2006, 7, 1326–1333. [Google Scholar] [CrossRef]
  150. Kim, H.-Y.; Skokos, E.A.; Myer, D.J.; Agaba, P.; Gonzalez, A.L. αVβ3 Integrin Regulation of Respiratory Burst in Fibrinogen Adherent Human Neutrophils. Cell. Mol. Bioeng. 2014, 7, 231–242. [Google Scholar] [CrossRef]
  151. Rossaint, J.; Herter, J.M.; Van Aken, H.; Napirei, M.; Oring, Y.D.; Weber, C.; Soehnlein , O.; Zarbock, A. Synchronized integrin engagement and chemokine activation is crucial in neutrophil extracellular trap-mediated sterile inflammation. Blood 2014, 123, 2573–2584. [Google Scholar] [CrossRef] [PubMed]
  152. Ueda, T.; Rieu, P.; Brayer, J.; Amin Arnaout, M. Identification of the complement iC3b binding site in the β2 integrin CR3 (CD11b/CD18). Proc. Nati. Acad. Sci. USA 1994, 91, 10680–10684. [Google Scholar] [CrossRef] [PubMed]
  153. Lau, D.; Mollnau, H.; Eiserich, J.P.; Freeman, B.A.; Daiber, A.; Gehling, U.M.; Brümmer, J.; Rudolph, V.; Münzel, T.; Heitzer, T.; et al. Myeloperoxidase mediates neutrophil activation by association with CD11b/CD18 integrins. Proc. Natl. Acad. Sci. USA 2005, 102, 431–436. [Google Scholar] [CrossRef]
  154. Newton, R.A.; Hogg, N. The Human S100 Protein MRP-14 Is a Novel Activator of the β2 Integrin Mac-1 on Neutrophils. J. Immunol. Ref. 1998, 160, 1427–1435. [Google Scholar] [CrossRef]
  155. Monsel, A.; Lécart, S.; Roquilly, A.; Broquet, A.; Jacqueline, C.; Mirault, T.; Troude, T.; Fontaine-Aupart, M.-P.; Asehnoune, K. Analysis of Autofluorescence in Polymorphonuclear Neutrophils: A New Tool for Early Infection Diagnosis. PLoS ONE 2014, 9, e92564. [Google Scholar] [CrossRef]
  156. Naegele, M.; Tillack, K.; Reinhardt, S.; Schippling, S.; Martin, R.; Sospedra, M. Neutrophils in multiple sclerosis are characterized by a primed phenotype. J. Neuroimmunol. 2012, 242, 60–71. [Google Scholar] [CrossRef]
  157. Andersen, M.N.; Al-Karradi, S.N.H.; Kragstrup, T.W.; Hokland, M. Elimination of erroneous results in flow cytometry caused by antibody binding to Fc receptors on human monocytes and macrophages. Cytom. Part A 2016, 89, 1001–1009. [Google Scholar] [CrossRef]
  158. Smirnov, A.; Solga, M.D.; Lannigan, J.; Criss, A.K. Using Imaging Flow Cytometry to Quantify Neutrophil Phagocytosis. Neutrophil. Methods Mol. Biol. 2020, 2087, 127–140. [Google Scholar] [CrossRef]
  159. Han, Y.; Gu, Y.; Zhang, A.C.; Lo, Y.-H. Review: Imaging technologies for flow cytometry. Lab Chip 2016, 16, 4639–4647. [Google Scholar] [CrossRef]
  160. Eckert, R.E.; Neuder, L.E.; Park, J.; Adler, K.B.; Jones, S.L. Myristoylated Alanine-Rich C-Kinase Substrate (MARCKS) Protein Regulation of Human Neutrophil Migration. Am. J. Respir. Cell Mol. Biol. 2010, 42, 586–594. [Google Scholar] [CrossRef]
  161. Takahashi, M.; Nagao, T.; Matsuzaki, K.; Nishimura, T.; Minamitani, H. Photodynamically induced endothelial cell injury and neutrophil-like HL-60 adhesion. J. Photoscience 2002, 9, 518–520. [Google Scholar]
  162. Shimizu, Y.; Mobley, J.L.; Finkelstein, L.D.; Chan, A.S.H. A role for phosphatidylinositol 3-kinase in the regulation of beta 1 integrin activity by the CD2 antigen. J. Cell Biol. 1995, 131, 1867–1880. [Google Scholar] [CrossRef] [PubMed]
  163. Shrestha, D.; Jenei, A.; Nagy, P.; Vereb, G.; Szöllősi, J. Understanding FRET as a Research Tool for Cellular Studies. Int. J. Mol. Sci. 2015, 16, 6718–6756. [Google Scholar] [CrossRef] [PubMed]
  164. Liu, W.; Hsu, A.Y.; Wang, Y.; Lin, T.; Sun, H.; Pachter, J.S.; Groisman, A.; Imperioli, M.; Yungher, F.W.; Hu, L.; et al. Mitofusin-2 regulates leukocyte adhesion and β2 integrin activation. J. Leukoc. Biol. 2022, 111, 771–791. [Google Scholar] [CrossRef] [PubMed]
  165. Zarbock, A.; Lowell, C.A.; Ley, K. Syk signaling is necessary for E-selectin-induced LFA-1-ICAM-1 association and rolling but not arrest. Immunity 2007, 26, 773–783. [Google Scholar] [CrossRef] [PubMed]
  166. Morikis, V.A.; Chen, S.J.; Madigan, J.; Jo, M.H.; Werba, L.C.; Ha, T.; Simon, S.I. β2-Integrin Adhesive Bond Tension under Shear Stress Modulates Cytosolic Calcium Flux and Neutrophil Inflammatory Response. Cells 2022, 11, 2822. [Google Scholar] [CrossRef]
  167. Kaboord, B.; Perr, M. Isolation of Proteins and Protein Complexes by Immunoprecipitation. In 2D PAGE: Sample Preparation and Fractionation; Posch, A., Ed.; Humana Press: Totowa, NJ, USA, 2008; pp. 349–364. ISBN 978-1-60327-064-9. [Google Scholar]
  168. Evans, R.; Lellouch, A.C.; Svensson, L.; McDowall, A.; Hogg, N. The integrin LFA-1 signals through ZAP-70 to regulate expression of high-affinity LFA-1 on T lymphocytes. Blood 2011, 117, 3331–3342. [Google Scholar] [CrossRef]
  169. Kumar, S.; Xu, J.; Perkins, C.; Guo, F.; Snapper, S.; Finkelman, F.D.; Zheng, Y.; Filippi, M.-D. Cdc42 regulates neutrophil migration via crosstalk between WASp, CD11b, and microtubules. Blood 2012, 120, 3563–3574. [Google Scholar] [CrossRef]
  170. Xiong, Y.-M.; Chen, J.; Zhang, L. Modulation of CD11b/CD18 Adhesive Activity by Its Extracellular, Membrane-Proximal Regions. J. Immunol. 2003, 171, 1042–1050. [Google Scholar] [CrossRef]
  171. Wang, J.-X.; Bair, A.M.; King, S.L.; Shnayder, R.; Huang, Y.-F.; Shieh, C.-C.; Soberman, R.J.; Fuhlbrigge, R.C.; Nigrovic, P.A. Ly6G ligation blocks recruitment of neutrophils via a β2-integrin–dependent mechanism. Blood 2012, 120, 1489. [Google Scholar] [CrossRef]
  172. Piccardoni, P.; Manarini, S.; Federico, L.; Bagoly, Z.; Pecce, R.; Martelli, N.; Piccoli, A.; Totani, L.; Cerletti, C.; Evangelista, V. SRC-dependent outside-in signalling is a key step in the process of autoregulation of β2 integrins in polymorphonuclear cells. Biochem. J. 2004, 380, 57–65. [Google Scholar] [CrossRef] [PubMed]
  173. Lukácsi, S.; Nagy-Baló, Z.; Erdei, A.; Sándor, N.; Bajtay, Z. The role of CR3 (CD11b/CD18) and CR4 (CD11c/CD18) in complement-mediated phagocytosis and podosome formation by human phagocytes. Immunol. Lett. 2017, 189, 64–72. [Google Scholar] [CrossRef] [PubMed]
  174. Xu, S.; Wang, J.; Wang, J.-H.; Springer, T.A. Distinct recognition of complement iC3b by integrins α X β 2 and α M β2. Proc. Natl. Acad. Sci. USA 2017, 114, 3403–3408. [Google Scholar] [CrossRef] [PubMed]
  175. Fan, Z.; McArdle, S.; Marki, A.; Mikulski, Z.; Gutierrez, E.; Engelhardt, B.; Deutsch, U.; Ginsberg, M.; Groisman, A.; Ley, K. Neutrophil recruitment limited by high-affinity bent β2 integrin binding ligand in cis. Nat. Commun. 2016, 7, 12658. [Google Scholar] [CrossRef]
  176. Lämmermann, T.; Germain, R.N. The multiple faces of leukocyte interstitial migration. Semin. Immunopathol. 2014, 36, 227–251. [Google Scholar] [CrossRef]
  177. Brown, A.F. Neutrophil granulocytes: Adhesion and locomotion on collagen substrata and in collagen matrices. J. Cell Sci. 1982, 58, 455–467. [Google Scholar] [CrossRef]
  178. Friedl, P.; Weigelin, B. Interstitial leukocyte migration and immune function. Nat. Immunol. 2008, 9, 960–969. [Google Scholar] [CrossRef]
  179. Lämmermann, T.; Bader, B.L.; Monkley, S.J.; Worbs, T.; Wedlich-Söldner, R.; Hirsch, K.; Keller, M.; Förster, R.; Critchley, D.R.; Fässler, R.; et al. Rapid leukocyte migration by integrin-independent flowing and squeezing. Nature 2008, 453, 51–55. [Google Scholar] [CrossRef]
  180. Toyjanova, J.; Flores-Cortez, E.; Reichner, J.S.; Franck, C. Matrix Confinement Plays a Pivotal Role in Regulating Neutrophil-generated Tractions, Speed, and Integrin Utilization. J. Biol. Chem. 2015, 290, 3752–3763. [Google Scholar] [CrossRef]
  181. Mizgerd, J.P.; Horwitz, B.H.; Quillen, H.C.; Scott, M.L.; Doerschuk, C.M. Effects of CD18 Deficiency on the Emigration of Murine Neutrophils During Pneumonia. J. Immunol. 1999, 163, 995–999. [Google Scholar] [CrossRef]
  182. Palmer, C.S.; Kimmey, J.M. Neutrophil Recruitment in Pneumococcal Pneumonia. Front. Cell. Infect. Microbiol. 2022, 12, 574. [Google Scholar] [CrossRef] [PubMed]
  183. Doerschuk, C.M.; Tasaka, S.; Wang, Q. CD11/CD18-Dependent and -Independent Neutrophil Emigration in the Lungs: How Do Neutrophils Know Which Route to Take? Am. J. Respir. Cell Mol. Biol. 2000, 23, 133–136. [Google Scholar] [CrossRef] [PubMed]
  184. Mileski, W.; Harlan, J.; Rice, C.; Winn, R. Streptococcus pneumoniae-stimulated macrophages induce neutrophils to emigrate by a CD18-independent mechanism of adherence. Circ. Shock 1990, 31, 259–267. [Google Scholar] [PubMed]
  185. Zenaro, E.; Pietronigro, E.; Della Bianca, V.; Piacentino, G.; Marongiu, L.; Budui, S.; Turano, E.; Rossi, B.; Angiari, S.; Dusi, S.; et al. Neutrophils promote Alzheimer’s disease–like pathology and cognitive decline via LFA-1 integrin. Nat. Med. 2015, 21, 880–886. [Google Scholar] [CrossRef]
  186. Orlova, V.V.; Choi, E.Y.; Xie, C.; Chavakis, E.; Bierhaus, A.; Ihanus, E.; Ballantyne, C.M.; Gahmberg, C.G.; Bianchi, M.E.; Nawroth, P.P.; et al. A novel pathway of HMGB1-mediated inflammatory cell recruitment that requires Mac-1-integrin. EMBO J. 2007, 26, 1129–1139. [Google Scholar] [CrossRef]
  187. Tak, T.; Karnam, G.; Boon, L.; Viveen, M.; Meyaard, L.; Koenderman, L.; Pillay, J.; Rygiel, T.P.; Bastian, O.W.; Coenjaerts, F.E. Neutrophil-mediated Suppression of Influenza-induced Pathology Requires CD11b/CD18 (MAC-1). Am. J. Respir. Cell Mol. Biol. 2018, 58, 492–499. [Google Scholar] [CrossRef]
  188. Kadioglu, A.; De Filippo, K.; Bangert, M.; Fernandes, V.E.; Richards, L.; Jones, K.; Andrew, P.W.; Hogg, N. The Integrins Mac-1 and α4β1 Perform Crucial Roles in Neutrophil and T Cell Recruitment to Lungs during Streptococcus pneumoniae Infection. J. Immunol. 2011, 186, 5907–5915. [Google Scholar] [CrossRef]
  189. Haist, M.; Ries, F.; Gunzer, M.; Bednarczyk, M.; Siegel, E.; Kuske, M.; Grabbe, S.; Radsak, M.; Bros, M.; Teschner, D. Neutrophil-Specific Knockdown of β2 Integrins Impairs Antifungal Effector Functions and Aggravates the Course of Invasive Pulmonal Aspergillosis. Front. Immunol. 2022, 13, 823121. [Google Scholar] [CrossRef]
  190. Kevil, C.G.; Hicks, M.J.; He, X.; Zhang, J.; Ballantyne, C.M.; Raman, C.; Schoeb, T.R.; Bullard, D.C. Loss of LFA-1, but not Mac-1, Protects MRL/MpJ-Faslpr Mice from Autoimmune Disease. Am. J. Pathol. 2004, 165, 609–616. [Google Scholar] [CrossRef]
  191. Mizgerd, B.J.P.; Kubo, H.; Kutkoski, G.J.; Bhagwan, S.D.; Scharffetter-kochanek, K.; Beaudet, A.L.; Doerschuk, C.M. Neutrophil Emigration in the Skin, Lungs, and Peritoneum: Different Requirements for CD11/CD18 Revealed by CD18-deficient Mice. J. Exp. Med. 1997, 186, 1357–1364. [Google Scholar] [CrossRef]
  192. Coxon, A.; Rieu, P.; Barkalow, F.J.; Askari, S.; Sharpe, A.H.; Von Andrian, U.H.; Arnaout, M.A.; Mayadas, T.N. A Novel Role for the β2 Integrin CD11b/CD18 in Neutrophil Apoptosis: A Homeostatic Mechanism in Inflammation. Immunity 1996, 5, 653–666. [Google Scholar] [CrossRef] [PubMed]
  193. Kranig, S.A.; Lajqi, T.; Tschada, R.; Braun, M.; Kuss, N.; Pöschl, J.; Hudalla, H. Leukocyte Infiltration of Cremaster Muscle in Mice Assessed by Intravital Microscopy. J. Vis. Exp. 2020, 2020, e60509. [Google Scholar] [CrossRef]
  194. Phillipson, M.; Heit, B.; Colarusso, P.; Liu, L.; Ballantyne, C.M.; Kubes, P. Intraluminal crawling of neutrophils to emigration sites: A molecularly distinct process from adhesion in the recruitment cascade. J. Exp. Med. 2006, 203, 2569–2575. [Google Scholar] [CrossRef] [PubMed]
  195. Lim, K.; Hyun, Y.-M.; Lambert-Emo, K.; Topham, D.J.; Kim, M. Visualization of integrin Mac-1 in vivo. J. Immunol. Methods 2015, 426, 120–127. [Google Scholar] [CrossRef] [PubMed]
  196. Margraf, A.; Ley, K.; Zarbock, A. Neutrophil Recruitment: From Model Systems to Tissue-Specific Patterns. Trends Immunol. 2019, 40, 613–634. [Google Scholar] [CrossRef]
  197. Behnen, M.; Leschczyk, C.; Möller, S.; Batel, T.; Klinger, M.; Solbach, W.; Laskay, T. Immobilized Immune Complexes Induce Neutrophil Extracellular Trap Release by Human Neutrophil Granulocytes via FcγRIIIB and Mac-1. J. Immunol. 2014, 193, 1954–1965. [Google Scholar] [CrossRef]
  198. Loike, J.D.; Cao, L.; Budhu, S.; Marcantonio, E.E.; El Khoury, J.; Hoffman, S.; Yednock, T.A.; Silverstein, S.C. Differential regulation of beta1 integrins by chemoattractants regulates neutrophil migration through fibrin. J. Cell Biol. 1999, 144, 1047–1056. [Google Scholar] [CrossRef]
  199. Lomakina, E.B.; Waugh, R.E. Adhesion Between Human Neutrophils and Immobilized Endothelial Ligand Vascular Cell Adhesion Molecule 1: Divalent Ion Effects. Biophys. J. 2009, 96, 276–284. [Google Scholar] [CrossRef]
  200. Senior, R.M.; Gresham, H.D.; Griffin, G.L.; Brown, E.J.; Chung, A.E. Entactin stimulates neutrophil adhesion and chemotaxis through interactions between its Arg-Gly-Asp (RGD) domain and the leukocyte response integrin. J. Clin. Investig. 1992, 90, 2251–2257. [Google Scholar] [CrossRef]
  201. Murphy, J.F.; Bordet, J.C.; Wyler, B.; Rissoan, M.C.; Chomarat, P.; Defrance, T.; Miossec, P.; McGregor, J.L. The vitronectin receptor (αvβ3) is implicated, in cooperation with P-selectin and platelet-activating factor, in the adhesion of monocytes to activated endothelial cells. Biochem. J. 1994, 304, 537–542. [Google Scholar] [CrossRef]
Table 1. Nomenclature for β2-integrins expressed on neutrophils.
Table 1. Nomenclature for β2-integrins expressed on neutrophils.
β2-IntegrinHeterodimerOther NamesLigands
CD11a/CD18αLβ2LFA-1ICAM-1, ICAM-2, LPS
CD11b/CD18αMβ2Mac-1, Complement receptor 3 (CR3)iC3b, fibrinogen, factor X, ICAM-1, LPS
CD11c/CD18αXβ2P150,95, Complement receptor 4 (CR4)Fibrinogen, iC3b, collagen, ICAM-1, LPS, β-glucan
CD11d/CD18αDβ2 ICAM-3, VCAM-1, fibronectin, vitronectin, fibrinogen
Table 2. Diseases and disorders where neutrophil β2-integrins have been identified as key players in disease.
Table 2. Diseases and disorders where neutrophil β2-integrins have been identified as key players in disease.
Disease or DisorderCategoryReferences
Antineutrophil cytoplasmic antibody (ANCA)-associated vasculitis (AAV)Autoimmune disease[19]
Aspergillus fumigatusFungal infections[20,21]
Atrial fibrillationCardiovascular disease[22]
Blood-brain barrier inflammationAcute illness[23,24]
Candida albicansFungal infections[25]
COPDChronic disease[26,27,28]
Interstitial lung disease (ILD)Chronic inflammation, autoimmune disease[29]
Ischemia-reperfusion injuryAcute injury/Sterile inflammation[30,31,32,33,34]
Leukocyte adhesion deficiency (LAD)Genetic disorder[35,36,37]
Myocardial InfarctionCardiovascular disease/Sterile inflammation[33,34,38,39]
Rheumatoid arthritisAutoimmune disease[40]
SARS-CoV-2Infectious disease[41,42]
SepsisAcute illness[43]
Sepsis-induced acute lung injuryAcute injury[44,45]
Solid organ transplant rejectionTransplant rejection[46]
Systemic lupus erythematosus (SLE)Autoimmune disease[12,40,47,48]
ThrombosisCardiovascular disease[49]
Transfusion-related acute lung injury (TRALI)Acute injury[50]
Trauma/Vascular injuryAcute injury[51,52,53]
Wiskott Aldrich syndromeGenetic disorder[54,55]
Table 3. Cell lines used in neutrophil β2-integrin research 1.
Table 3. Cell lines used in neutrophil β2-integrin research 1.
Cell LineRequires DifferentiationEndogenous β2-Integrin ExpressionLimitations/Drawbacks
HL60/PLB-985Yes—DMSO or retinoic acid (referred to as dHL60sαLβ2, differentiation required for αMβ2
  • Lack of specific and secretory granules, which limits upregulation of αMβ2 following stimulation
  • DMSO dHL60s have different IL-8R, signaling proteins, and α-actinin compared with humans.
  • Lower αMβ2 expression and dampened upregulation by fMLP and LTB4 compared with humans
K562NoNo
  • Requires stable transfection for αMβ2 expression
HoxB8Yes—GM-CSF or engraftment into miceYes—αLβ2 and αMβ2
  • Additional cytokines are needed to achieve “mature” neutrophilic cells.
  • Lower levels of ROS production compared with murine neutrophils due to decrease gp91phox expression.
  • Lower chemotactic responses
  • Not suitable for degranulation experiments
1 see the article text for references relevant to Table 3.
Table 5. Evidence for β2-integrin dependence in murine models in vivo.
Table 5. Evidence for β2-integrin dependence in murine models in vivo.
Disease Modelβ2-Integrin DependentReferences
Alzheimer’s diseaseYes[185]
Atrial fibrosisYes[22]
HMGB1-induced peritonitisMac-1: Yes
LFA-1: No
[186]
InfluenzaNo[187]
LTB4-induced intestinal
transepithelial migration
Yes[187]
Pneumonia
S. pneumoniaMac-1: yes
LFA-1: no
[188]
[181]
P. aeruginosaYes[127,181]
E. coli LPSYes[181]
Pulmonary aspergillosisYes[20,189]
SLE-induced glomerular diseaseMac-1: No
LFA-1: Yes
[190]
Thioglycollate peritonitisMac-1: No
LFA-1: Yes
[186,191,192]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Conley, H.E.; Sheats, M.K. Targeting Neutrophil β2-Integrins: A Review of Relevant Resources, Tools, and Methods. Biomolecules 2023, 13, 892. https://doi.org/10.3390/biom13060892

AMA Style

Conley HE, Sheats MK. Targeting Neutrophil β2-Integrins: A Review of Relevant Resources, Tools, and Methods. Biomolecules. 2023; 13(6):892. https://doi.org/10.3390/biom13060892

Chicago/Turabian Style

Conley, Haleigh E., and M. Katie Sheats. 2023. "Targeting Neutrophil β2-Integrins: A Review of Relevant Resources, Tools, and Methods" Biomolecules 13, no. 6: 892. https://doi.org/10.3390/biom13060892

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop