Next Article in Journal
Surprising Structural and Functional Properties of Favism Erythrocytes Are Linked to Special Metabolic Regulation: A Cell Aging Study
Next Article in Special Issue
The Epidermal Keratinocyte as a Therapeutic Target for Management of Diabetic Wounds
Previous Article in Journal
The Role of Long Noncoding RNA (lncRNAs) Biomarkers in Renal Cell Carcinoma
Previous Article in Special Issue
Evaluating Thera-101 as a Low-Volume Resuscitation Fluid in a Model of Polytrauma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Acute Inflammation in Tissue Healing

by
Amro M. Soliman
1 and
Daniel R. Barreda
1,2,*
1
Department of Biological Sciences, University of Alberta, Edmonton, AB T6G 2R3, Canada
2
Department of Agricultural, Food and Nutritional Science, University of Alberta, Edmonton, AB T6G 2R3, Canada
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(1), 641; https://doi.org/10.3390/ijms24010641
Submission received: 11 November 2022 / Revised: 19 December 2022 / Accepted: 28 December 2022 / Published: 30 December 2022
(This article belongs to the Special Issue Wound Repair and Regeneration 2022)

Abstract

:
There are well-established links between acute inflammation and successful tissue repair across evolution. Innate immune reactions contribute significantly to pathogen clearance and activation of subsequent reparative events. A network of molecular and cellular regulators supports antimicrobial and tissue repair functions throughout the healing process. A delicate balance must be achieved between protection and the potential for collateral tissue damage associated with overt inflammation. In this review, we summarize the contributions of key cellular and molecular components to the acute inflammatory process and the effective and timely transition toward activation of tissue repair mechanisms. We further discuss how the disruption of inflammatory responses ultimately results in chronic non-healing injuries.

1. Introduction

Throughout millions of years of evolution, multicellular organisms have adapted complex systems for recognizing and repairing injured tissues [1]. Reparative responses share common pathways and characteristics, where the inflammatory response is considered a cornerstone for effective tissue repair [2]. Upon injury, inflammatory cascades are initiated to clear pathogens and regulate subsequent healing events. Still, tight regulation of acute inflammation is required to avoid its excessive perturbations, which ultimately results in defective and delayed healing. We previously highlighted some of the molecular and cellular mechanisms regulating the induction and resolution of inflammation during cutaneous infection [3]. These mechanisms further influence other parts of the healing process and potentially determine the outcome.
Tissue repair is a complex biological reaction to an injury involving interactions between various immune and connective tissue cells. Together, these cells and several humoral factors accomplish sequential phases, comprising hemostasis, inflammation and proliferation, to restore damaged tissues [4]. Subsequent to tissue damage, there is a constriction of injured blood vessels followed by platelet activation and clot formation to stop bleeding [5]. Fibrin threads act as a scaffold for infiltrating leukocytes. Neutrophils are among the early cells relocating to the injury site, representing the first line of defense against infection [6]. Monocytes follow neutrophils, where they mature to macrophages [7]. The cellular migration is triggered mainly via activation of an acute inflammatory program that involves expression of several cytokines and chemokines [8]. Following the eradication of infection and tissue debris, anti-inflammatory mediators and growth factors are released to suppress inflammation and initiate the proliferative phase [9]. Several tissue repair events, including angiogenesis, granulation tissue formation and re-epithelialization, are initiated.
The newly formed granulation tissue and epidermis require an adequate supply of nutrients and oxygen to maintain homeostasis and to promote further healing. This is achieved by developing new blood vessels at the wound area. The process is known as neovascularization or angiogenesis, accomplished via activating local endothelial cells (ECs) lining the inner surface of neighboring blood vessels [10]. In response to hypoxia-responsive growth factors (e.g., VEGF), ECs migrate, proliferate and form new cell-to-cell junctions to develop new capillaries branching out from existing blood vessels. Angiogenesis is regulated by angiogenic molecules other than VEGF, such as platelet-derived growth factor (PDGF), fibroblast growth factor (FGF) and angiopoietins [11]. ECs show heterogeneity in response to these molecules by functioning as either lead tip or trailing stalk cells. Lead tip cells migrate towards angiogenic factors in response to positive and negative regulators. On the other hand, stalk cells preserve the structure of existing blood vessels [12].
During the proliferation phase, new connective tissue is formed concurrently with neovascularization. Fibroblast is the key cell accountable for constructing granulation tissue to fill in the wound gap. In response to various signaling molecules that are released from tissue-resident macrophages, platelets, keratinocytes and ECs (e.g., transforming growth factor (TGF)-β, EGF and PDGF) [13,14,15], fibroblasts proliferate, migrate and become pro-fibrotic, depositing ECM proteins [16]. Prior to laying ECM proteins by fibroblasts, they obliterate provisional matrix by secreting MMPs, to be substituted by granulation tissue rich in collagen, fibronectin, glycoproteins and proteoglycans [17]. Fibroblasts in wound repair demonstrate functional diversity, stemming from their contribution to many other healing events [18]. Fibroblasts exhibit heterogeneity depending on their activation status and throughout different developmental stages [19], resulting in significant phenotypic differences. These phenotypes mediate varying functions in wound repair, including re-epithelialization [20], secretion of growth factors, immunomodulation, ECM synthesis and organization [21].Keratinocytes differentiate and migrate from wound edges toward the center to cover wound surface. This is achieved by weakening of cell–cell connections and adhesion to basement membrane [22]. Cellular migration continues until keratinocytes from opposing edges come in contact, forming the basal layer that develops new adhesions with the underlying matrix [23]. Suprabasal keratinocytes proliferate to provide multiple overlying layers of keratinocytes [24]. Re-epithelialization is regulated via several cytokines and growth factors as well as crosstalk between keratinocytes and inflammatory cells (PMN and macrophages), fibroblasts and ECs [25]. For instance, keratinocytes were found to activate fibroblasts to release growth factors, thus promoting keratinocyte proliferation [15]. Indeed, regulation of proliferation phase of tissue repair is orchestrated by various cytokines and growth factors released by inflammatory cells such as macrophages [26]. It is further dependent on efficient clearance of pathogens and proper resolution of inflammation [27]. This is clearly observed in disrupted tissue repair associated with conditions of impaired immune responses and prolonged inflammation that are instigated by, for example, diabetes, aging and malnutrition [28,29]. This review discusses the significant contribution of acute inflammatory reaction’s various components and effectors to tissue repair. Recognizing these regulatory factors is critical for understanding physiology of the repair process and pathophysiology of chronic non-healing injuries commonly associated with their disruption.

2. Induction of Inflammation Phase of Tissue Repair: Innate Immune Responses

Following an injury, danger/damage-associated molecular patterns (DAMPs) [30], generated by necrotic cells, as well as pathogen-associated molecular patterns (PAMPs) [31], including conserved motifs of invading pathogens, are recognized by innate receptors on tissue-resident cells to trigger an acute inflammatory reaction [17] (Figure 1). As a result, various inflammatory mediators are released to promote leukocyte recruitment and regulate immune responses at the injury site [32].

2.1. Immune System Perception of Injury: The Role of DAMPs and PAMPs

DAMPs are either passively or actively released by injured host cells [33,34]. They include patterns that are usually isolated inside cells with limited extracellular exposure. DNA (genomic or mitochondrial), ATP and other peptides are released into the extracellular space following cell death or lysis [35,36,37,38]. These patterns provide self-injury detection strategies for the host to activate inflammatory responses in the case of sterile injuries or wounds with restricted PAMPs. Mechanistically, DAMPs function by directly binding to pattern recognition receptors (PPRs) on resident cells or indirectly by modifying extracellular matrix (ECM) molecules such as heparan sulfate, fibrillary protein and collagen to possess proinflammatory-stimulating properties [39,40,41]. In addition to DAMPs, hydrogen peroxide (H2O2), chemokines and other lipid mediators are released by damaged cells to deliver signals promoting innate immune functions such as leukocyte recruitment. For example, H2O2 provokes neutrophil migration toward the injury site for rapid pathogen clearance [42]. On the other hand, PAMPs, also known as microbe-associated molecular patterns (MAMPs), represent parts of different invading microbes (e.g., viruses, bacteria, parasites and fungi). They embrace LPS, microbial lipoproteins, β-glucan and double-stranded RNA [43,44,45,46,47]. Similar to DAMPs, PAMPs can act as ligands for several PPRs to launch acute inflammatory programs.

2.2. Activation of PPRs and Downstream Inflammatory Pathways

The injury-recognition system markedly relies on innate PPRs located on tissue-resident cells. PRRs include several types, such as Toll-like receptors (TLRs), nucleotide-binding oligomerization domain-like receptors (NLRs), retinoic acid-inducible gene I like receptors (RLRs) and C-type lectin receptors (CLRs) [48]. TLRs were found to play a vital role in initiating the inflammatory phase of tissue repair [49] via specifically binding a variety of ligands. For instance, TLR1, TLR2 and TLR5 can detect bacterial peptidoglycans and flagellin, while TLR3, TLR7 and TLR8 recognize single and double-stranded RNA of viruses [50]. Activated TLRs trigger nuclear factor-kB (NF-κB) and mitogen-activated protein kinase (MAPK) pathways via adaptor proteins such as TRAM, TRIF, MyD88 and TIRAP/Mal, thus upregulating the expression of proinflammatory cytokines, such as IL-1β and TNF-α in addition to other chemokines, lipid mediators and adhesion molecules [51,52]. Interestingly, IL-1β and TNF-α provide a positive feedback loop by directly activating NF-κB-facilitated gene expression, thereby intensifying inflammatory responses [53]. Other transcription-independent pathways are activated early at the injury site via, for example, Ca2+ influx and reactive oxygen species (ROS) to compensate for the delay in the induction of transcription machinery [26]. Intracellular Ca2+ levels increased substantially within a few minutes at wound edges and later at the center after an injury [54]. All these pathways contribute to initiating an acute inflammatory state.

2.3. Inflammatory Cytokines and Mediators

Cytokines are small proteins (~10 kDa) characterized by having the amino acid, cysteine in their structure. The two most important cytokines involved in tissue repair are (1) CC cytokines which have two adjacent cysteines and (2) CXC cytokines which contain two cysteines separated by another amino acid. Cytokine production is a complex biological process regulated by several activators and inhibitors depending on the stage of inflammation and other environmental factors [55]. Cytokines are essential for the induction, propagation and resolution of the inflammatory phase of tissue repair [56]. Moreover, they promote cellular recruitment and regulate their development, proliferation and functions during the healing process, as shown in Table 1. Notably, most tissue-resident cells, including parenchymal cells, fibroblasts, ECs, and immune cells, can produce various cytokines in response to DAMPs and PAMPs. Furthermore, recruited leukocytes accentuate the release of these cytokines under a proinflammatory condition [57]. The pleiotropic properties of inflammatory cytokines allow them to exert a wide range of functions. For instance, despite its critical role in the induction of acute inflammation and leukocyte chemotaxis, IL-6 was also found to promote resolution [58,59].
Other crucial mediators of inflammation were shown to influence the healing process. For instance, upon injury, mast cells are triggered to release histamine, which in addition to increasing vascular permeability, it stimulates proliferation of keratinocytes [60] and fibroblasts to activate collagen synthesis [61]. Moreover, histamine is essential for platelet and integrin activation [62]. Prostaglandin (PG)E2, another mediator released early during wound healing, significantly modulates inflammatory and reparative responses [63]. PGE2 augments VEGF expression, regulating angiogenesis [64], and activates M2 phenotype of macrophages, promoting tissue repair [65]. This was further supported by impaired healing and excessive scar formation induced by PGE2 inhibition [66]. On the other hand, thromboxane (TX)A2, released by activated platelets immediately upon injury, contributes to platelet aggregation necessary for hemostasis [67]. It further induces the synthesis of IL6 and PGE2 [68] and promotes angiogenesis by enhancing EC migration [69]. Being a critical chemoattractant, leukotriene B4 (LTB4) enhances recruitment of various immune cells to the injury site [70]. However, uncontrolled production and release of LTB4 delayed resolution of neutrophils in diabetic mice [71].
Table 1. Cytokines involved in tissue repair and their potential biological functions.
Table 1. Cytokines involved in tissue repair and their potential biological functions.
CytokineReceptorSourceFunctions
TNF-αTNFR1 (p55) and TNFR2 (p75)PMN, macrophages and mast cells
Induction of acute inflammation and cellular recruitment [72,73]
Increases synthesis of adhesion molecules to augment PMN recruitment [74]
Promotes angiogenesis [75]
Enhances keratinocyte proliferation and expression of adhesion molecules [76]
Mast cell-released TNF-α activates DCs migration and maturation [77]
Acts as a mitogen for fibroblasts [78]
Stimulates expression of PAFR and EGF receptors, enhancing cellular migration and proliferation [79]
IL-1βIL-1R1, IL-1R2 and IL-1RAcP (IL-1R3)keratinocytes, PMN and macrophages
Induction of acute inflammation and cellular recruitment [72,73]
Augments producing other proinflammatory cytokines, including TNF-α and IL-6 [80]
Increases fibroblast-secreted KGF and FGF-7 to promote keratinocyte migration and proliferation [56,81,82]
Triggers skin stem cell proliferation and activates gamma delta (γδ) T cells [83]
Stimulates myofibroblasts to produce proteases, degrading damaged ECM as well as facilitating fibroblast migration and proliferation in response to mitogens [84]
Promotes in vitro fibroblast migration and proliferation [85]
Increases the expression of VEGF (proangiogenic factors) and MMP-1 (enhances ECM degradation) [84]
Decreases expression of ECM proteins (e.g., collagen) and myofibroblast differentiation [84]
CXCL8CXCR1Platelets, PMN and macrophages
A potent chemoattractant of neutrophils [86,87]
Upregulates integrins and PMN-endothelium interactions to facilitate diapedesis [88]
Enhances antimicrobial mechanisms of PMN, such as ROS production and release of neutrophilic granules [89]
IL-6gp130 and IL-6RMyeloid cells, lymphocytes and fibroblasts
Induction of acute inflammation and cellular recruitment [72,73,90,91]
Induces Th2 and Th17 differentiation in CD4+ T-cells [92]
IL-6-stimulated Th2 cell release IL-4 and IL13 to activate polarization of M2 macrophage [93]
Promotes TGF-β expression [94] and re-epithelialization [95]
Enhances fibroblast proliferation, activation and migration [96]
Augments wound closure and granulation tissue formation in glucocorticoid-induced immunosuppressed mice [97]
Stimulates fibroblast differentiation to myofibroblast [98]
Activates fibroblasts, macrophages and keratinocytes to secrete VEGF, thus promoting angiogenesis [99]
IFN-γIFNGR1 and IFNGR2Natural killer cells, plasmacytoid DCs and T cells
Antiviral activities [74]
Activates macrophage to produce proinflammatory cytokines and enhances phagocytosis [100]
Regulates differentiation of CD4+ T cells into Th1 effectors [100]
Inhibits angiogenesis and collagen deposition via downregulating TGF-β-mediated signaling pathways [101,102]
IL-10IL-10RMacrophages, DCs, PMN, mast cells and T cells
Induction of anti-inflammatory responses and resolution of inflammation [103]
Inhibits expression of proinflammatory cytokines, chemokines and adhesion molecules in macrophages and neutrophils [104]
Suppresses NO and ROS production [104]
Promotes migration and invasion of fibroblasts [103]
Induces and maintains production of hyaluronic acid [105]
Protects against excessive collagen deposition and reduces scar formation [106]
TGF-βtype II TGF-β receptorMacrophages, keratinocytes, fibroblasts and platelets
Antagonize PMN chemoattractants (e.g., IL-8) and suppresses migration of inflammatory cells to the injury site [107,108]
Enhances the expression of ECM components such as collagen and fibronectin by fibroblasts [56] and inhibits various MMPs [109]
Promotes angiogenic activities of endothelial progenitor cells [110]
Augments keratinocyte migration and overall re-epithelialization [111]
Stimulate transformation of fibroblasts to myofibroblasts through the acquisition of αSMA via SMAD-dependent and independent transcriptional activity [112,113]
TNF-α: tumor necrosis factor-alpha; TNFR: tumor necrosis factor receptor; PMN: polymorphonuclear leukocytes; IL: interleukin; KGF: keratinocyte growth factor; FGF: fibroblast growth factor; CXCL8: C-X-C motif chemokine ligand 8; TGF-β: transforming growth factor-beta; IFNγ: interferon-gamma; DCs: dendritic cells; NO: nitric oxide; ROS: reactive oxygen species; ECM: extracellular matrix; MMPs: matrix metalloproteinase; PAFR: platelet activating factor receptor; VEGF: vascular endothelial growth factor; αSMA: alpha smooth muscle actin.

2.4. Cellular Recruitment to Injury Site

A number of CC and CXC cytokines act as chemoattractant proteins, identified as chemokines, regulating the migration of several immune and non-immune cells that are critical for the repair process. More than fifty chemokines and eighteen chemokine receptors have been characterized in humans and mice [114]. Production and diffusion of chemokines have to be precisely controlled to tailor their availability, thus accurately directing migrating cells [115]. Unfortunately, mechanisms regulating chemokine synthesis and diffusibility are still largely uncharacterized. It is worth mentioning that particular chemokines are constitutively produced and released under normal conditions to maintain tissue homeostasis via regulating basal cell functions and trafficking [116,117]. Though, in response to an injury, chemokines are expressed at higher levels during an acute inflammatory reaction to execute necessary immune responses. CXC chemokines containing glutamate–leucine–arginine (ELR) motifs such as C-X-C motif chemokine ligand 8 (CXCL8), also known as IL-8, are more specialized in polymorphonuclear leukocyte (PMN) recruitment [118,119]. Meanwhile, other CC cytokines, including C-C motif ligand 1 (CCL1) act as monocyte [120] and lymphocyte [121] attractants. The expression of these chemokines must be firmly regulated during tissue repair to avoid dysregulation of inflammatory responses. Persistent uncontrolled expression of particular chemoattractants results in the development of a variety of pathological conditions [115,122,123,124,125,126]. Upon their release, chemokines bind to glycosaminoglycans on ECs of blood vessels to be presented to circulating immune cells [127]. Leukocytes bind to these chemokines via their corresponding G protein-coupled receptors (GPCRs), resulting in the extravasation of these cells and migration toward the injury site [128,129]. For instance, C-X-C motif chemokine receptor 1 (CXCR1) and CXCR2 on PMN bind to CXCL8, activating downstream signaling pathways and promoting neutrophil recruitment [130]. Intriguingly, as reviewed by Ridiandries et al. [95], several chemokines were shown to contribute not only to inflammatory cell recruitment but also to proliferation and remodeling phases of tissue repair. For instance, CXCL1 and CXCL7 were involved in angiogenesis [131], while CXCL12 promotes differentiation of stem cells into fibroblasts and ECs, enhancing granulation tissue formation [132,133].
Leukocyte extravasation is achieved through several steps involving adhesion molecules (e.g., selectins and integrins), chemokines and interactions with ECs. Selectins (E-, P- and L-selectins), type I transmembrane proteins, are principally responsible for the initial tethering and adhesive interactions between ECs and circulating leukocytes [134]. Therefore, these adhesion molecules, such as E-selectin, were found to be upregulated by proinflammatory cytokines [135]. Selectins bind to carbohydrate-based ligands such as P-selectin glycoprotein ligand-1 (PSGL-1), generally expressed on leukocyte microvilli, to secure these cells to ECs at the injury site, where they mediate most tethering and rolling [136]. Integrins, on the other hand, are expressed on leukocytes, where they get activated by proinflammatory cytokines to induce cellular adherence to counter-receptors, such as intercellular adhesion molecules (ICAMs), enhancing adhesion of circulating immune cells to endothelium [137,138]. During leukocyte recruitment, various integrins are activated during different steps of transendothelial migration, where they were also found to support cellular arrest and rolling [139]. The integrin-counter receptor contact stimulates signaling cascades in captured leukocytes to achieve an intermediate-affinity conformational change, resulting in the “slow rolling” of these cells [136,140]. Pathways involved in integrin activation of leukocyte rolling, including inside-out and outside-in signaling mechanisms, are reviewed by McEver et al. [140]. Macrophage-1 antigen (Mac-1), a part of integrin cell surface receptors, was reported to be a crucial mediator of monocyte rolling with the help of P-selectin on low-shear stress ECM substrates [141]. Integrins can also stimulate cytokine secretion directly in neutrophils and macrophages [142]. Leukocytes roll along the surface of the endothelium to sense glycosaminoglycan-bound chemokines. Activation of chemokine receptors on leukocytes results in confirmational changes and leukocyte adhesion cascade (diapedesis) [139,143,144], where inflammatory cells crawl through endothelial junctions or weak regions of the basement membrane [145]. In a typical repair process, there is usually an early expression of neutrophil chemoattractants, resulting in rapid recruitment of PMN to the injury site. This is followed by subsequent waves of infiltrating monocytes and lymphocytes triggered by other chemokines [146].
In addition to the conventional pathway of inflammation-induced upregulation of chemokines gene expression, another recently reported alternative mechanism for leukocyte recruitment could be attained by mast cells. These cells contain granules filled with proinflammatory cytokines, vascular permeability and vasodilation factors, as well as proteases. Immediately after an injury, these molecules are released to enhance migration of immune cells to the injury site [147]. Interestingly, mice with a deficiency in mast cell proteases showed remarkably impaired neutrophil recruitment [148]. These findings do not necessarily abolish the significant role of chemokine-mediated PMN migration; nevertheless, it provides an early pathway for neutrophils to exist rapidly at the injury site for early pathogen clearance and compensate for a possible delay in the activation of transcription machinery.

3. Role of Inflammatory Cells during Tissue Repair

3.1. Neutrophils

Among immune cells involved in the repair process, neutrophils are considered the “first responders” since they are swiftly recruited [149,150], constituting approximately 50% of all cells at the injury site within 24 h after injury [151]. In addition to the potent CXCL8, other cytokines, such as CXCL4 and CCL3/4 promote PMN migration [151]. Notably, neutrophils are not commonly detected in healthy skin; instead, they remain in the bone marrow and bloodstream [152], ready to be drafted, as discussed in the previous section. Recruited neutrophils can augment additional PMN infiltration by releasing several chemoattractant factors [150,153,154]. The primary function of neutrophils at the injury site is to combat invading pathogens via various antimicrobial responses, including phagocytosis, toxic granules, oxidative burst and neutrophil extracellular traps (chromatin filaments released extracellularly to immobilize and eliminate microbes, known as NETs) [149,155]. Still, a critical balance must be maintained between phagocytes’ protective functions and their possible contributions to prolonged and exacerbated inflammation [156]. This equilibrium ensures the eradication of infection while minimizing collateral tissue damage. Several studies suggested that the prolonged existence of neutrophils at the injury site was detrimental to proper tissue repair [6,157]. This was attributed to PMNs-derived proteases degrading ECM and being allied with a deleterious oxidative burst [158,159]. Recently, neutrophils were also found to induce genomic instability via ROS-independent pathway involving the release of microparticles containing proinflammatory microRNAs (miR-23a and miR-155) in patients with inflammatory bowel disease (IBD) [160]. These miRNAs promoted the accumulation of double-strand breaks (DSBs) by inhibiting homologous recombination (HR), resulting in impeding inflammation resolution and overall intestinal healing [160].
Neutrophils can engulf bacteria and tissue debris through phagocytosis. Although the process is similar to that of macrophages, PMN possess distinctive phagocytic receptors [161]. As a result, PMN could handle antigens differently, where they can be opsonized or not. Fc receptors such as CD16, CD32 and CD64 recognize pathogens and then activate downstream Src and Rho-GTPases phagocytosis pathway. The result is an extension of the cell membrane to surround the antigen, forming phagocytic cups that get sealed to form phagosomes [162]. Additionally, neutrophils are characterized by having distinct granules containing bactericidal agents. These granules either fuse with phagosomes to destroy the pathogen intracellularly [163] or undergo exocytosis to combat microbes extracellularly [152]. Antimicrobial agents of these granules include myeloperoxidase, lysozyme, matrix metalloproteases (MMPs), lactoferrin and proteases (e.g., elastase and capthepsin G) [6]. The utilization of proteases by neutrophils is not limited to their anti-bacterial activity. Proteases are likewise crucial for neutrophil extravasation via degrading ECM elements and basement membrane of ECs directly or indirectly by activating MMPs [164]. Despite their importance for PMN migration and bactericidal actions, an unrestrained increase in proteoses induces extensive tissue damage, ensuing impaired healing and chronic wounds. This is on top compounded by proteolytic enzymes-induced obliteration of growth factors, newly formed blood vessels and granulation tissue [165]. Among these proteoses, elastase is released in response to bacterial infection in either free or membrane-bound form, and it was found to induce ECM destruction and direct epithelial damage [166]. Soluble elastase causes damage to areas surrounding neutrophils at degranulation site, while membrane-bound elastase can travel distally, resisting inhibition by anti-proteinases [167]. Elastase further contributes to neutrophil disintegration and NETs release by translocating to the nucleus and degrading chromatin through splitting histones [155]. Elastase can upregulate chemokines (e.g., IL-8) and other proteinases (MMP-9), instigating a vicious cycle of neutrophile recruitment and inflammation-associated collateral tissue injury [168]. Moreover, elastase degrades T-cell receptors and blocks antigen presentation, thus impairing lymphocytic functions [169].
Recent experimental evidence suggested parallel immunomodulatory functions of neutrophils during tissue healing in addition to their bactericidal actions. This was observed in mice with myocardial infarction (a sterile injury model), where researchers have characterized N2 neutrophils to play a potential role in restoring injured tissue irrespective of their antimicrobial functions [170]. Mechanistically, neutrophils modulate macrophage phenotype from a proinflammatory to anti-inflammatory/reparative state following engulfing apoptotic PMN by these macrophages, a process is known as efferocytosis [171,172]. Modulated macrophages release proresolution cytokines (e.g., IL-10) and growth factors such as transforming growth factor-beta (TGF-β) to control inflammation and initiate healing of damaged tissue [171]. A genetically modified reduction in PMN recruitment (CXCR2−/−) in injured mice resulted in delayed re-epithelialization at wound sites [173]. Likewise, aging-induced delayed wound healing was postulated to be instigated by the downregulation of neutrophil numbers or functions in mice [174] and humans [175,176]. This was attributed to impaired neutrophil-tempted pathogen clearance and, therefore, late resolution of inflammation. Conflicting data showed an accelerated re-epithelialization during neutrophil depletion [177]. Even though differences were observed at the level of epidermis development, no significant changes were evidenced in the dermis in terms of collagen deposition [177]. Still, further research is encouraged to characterize other possible immunomodulatory functions of PMN during tissue repair.

3.2. Macrophages

Macrophages play a critical role in tissue repair stemming from influencing both the inflammatory and proliferative phases. Macrophages’ contributions to immunomodulation, resolution of inflammation and tissue healing have been well-studied [178]. Macrophage numbers increase gradually at the injury site and peak 48–72 h after injury [179]. Influenced by chemokines such as CCL1 and CXCL12 [180], monocytes migrate to the injury site from bone marrow and adjacent blood vessels. Additionally, recruited macrophages can amplify the relocation of additional monocytes via releasing monocyte chemoattractant protein (MCP)-1 [181].
Several macrophage phenotypes were characterized during tissue repair [182,183,184]. It is worth mentioning that these phenotypes are not distinctively represented by particular macrophage subsets or a subject of on/off switch but rather a dynamic continuation of macrophage polarization based on environmental stimuli and interplay with other cells [185,186]. For instance, during the early phases of tissue repair, a classically activated macrophage phenotype, also known as M1, was shown to induce proinflammatory and bactericidal activities via expressing IL-1β and TNF-α in addition to mediating phagocytosis [187]. Later during the repair process, macrophages transition to becoming alternatively activated (M2) macrophages to suppress inflammation and promote the healing of damaged tissues [182]. Interestingly, recent reports indicated that M2 phenotype activation has expanded to involve other phenotypes triggered by various stimuli, such as M2a, M2b, M2c and M2d [184]. For example, M2a is activated by IL-4 and IL-13, while exposure to IL-10 and glucocorticoids stimulates M2c phenotype [184]. These M2 phenotypes largely intercede in anti-inflammatory, proresolution and healing roles [188]. Notably, macrophage phenotypes are not limited to the previously mentioned categories. There are likely several other phenotypes that are continuously activated depending on the differentiation stage, type and duration of stimulus as well as overall biochemical milieu [186,189].
The expanding literature supports the crucial role of macrophages in normal tissue repair. For instance, depletion of macrophages in wounds of murine models was associated with delayed healing induced by impaired angiogenesis, collagen synthesis and growth factors expression [190,191,192], indicating a significant engagement of macrophages in various repair events. Table 2 summarizes the potential functions and contributions of macrophage phenotypes during tissue healing.

3.3. Dendritic (DCs) and Langerhans Cells (LCs)

Since the discovery of DCs by Ralph Steinman in 1973, their role in immune responses and tissue homeostasis has been widely examined [209,210,211]. DCs can generally be categorized into tissue-resident and circulating DCs, also known as plasmacytoid DCs [212]. Tissue-resident DCs reside in tissues for immune surveillance, while plasmacytoid DCs are commonly absent in healthy tissues and are recruited following an injury [213,214]. The primary mission of DCs is to function as antigen-presenting cells that deliver antigens to T cells (CD8+ T cytotoxic and CD4+ T helper cells), establishing a bridge between innate and adaptive immunity [215]. The role of DCs in tissue repair was investigated mainly in murine models of wound healing, where DCs were found to be crucial for normal reparative responses. For instance, depletion of the early infiltrating plasmacytoid DCs after injuries considerably impaired the expression of proinflammatory cytokines and delayed re-epithelialization during wound healing in mice [213]. Moreover, a significant reduction in wound closure rate and granulation tissue deposition was observed in transgenic mice with depleted DCs [216]. Mechanistically, researchers suggested DCs to promote fibroblast proliferation and collagen synthesis in burn wounds. Recent evidence indicated a cross-talk between DCs and epithelial cells to maintain tissue homeostasis and promote tissue repair in the cornea. During corneal wound healing, DCs were shown to migrate with epithelial sheets to cover the wound surface [217]. On the other hand, the depletion of DCs resulted in an interruption of wound closure. Moreover, it significantly reduced epithelial cells-expressed CXCL10, IL-1β and thymic stromal lymphopoietin [217].
In the skin, DCs are identified as Langerhans cells (LCs), constituting approximately 2% of the epidermal cells [218]. Following an injury, LCs, similar to DCs, conduct antigen presentation after phagocytosing pathogens [218]. LCs can extend their dendrites through cellular tight junctions or completely reposition to reach microbes [219], inclined by cytokines that are primarily secreted by adjacent keratinocytes (e.g., MCP-1) [220]. Upon engulfing antigens, LCs translocate from epidermis to dermis layer by downregulating e-cadherin expression and utilizing the MMPs-induced degradation of basement membrane and ECM [221,222]. Guided by chemokines such as CXCL12, LCs migrate to draining lymph nodes, activating cell-mediated adaptive immune responses [223]. LCs can also keep tissue homeostasis by enhancing the activation and proliferation of T regulatory cells (Treg) [224,225]. Moreover, a significant role of LCs in diabetic wound healing was reported, where high numbers of LCs in diabetic foot ulcers were associated with a better healing outcome [226].

3.4. Mast Cells

Mast cells are specialized secretory cells differentiating from their precursors that migrate from bone marrow to perivascular regions of various connective tissues [227]. Since their discovery, mast cells have been recognized for their essential role in allergic reactions and combating parasitic infestations. However, many investigators recently became interested in studying their contributions to tissue repair. The growing evidence indicates mast cells to be critical for all phases of the repair process; hemostasis, inflammation and proliferation, as shown in Table 3. This is achieved primarily via cross-talk between mast cells and several other cells involved in tissue healing, modulating and triggering various activities [228]. Mast cell numbers increase at the injury site to reach about fivefold their original numbers, which was suggested to be a chemokine-mediated migration rather than cellular proliferation [151]. Despite the lack of studies examining chemokines regulating the migration of mast cells, keratinocytes-secreted MCP-1 was proposed to facilitate mast cell recruitment [151,229]. Mast cells are activated subsequent to binding of their receptors to a variety of ligands [230]. Upon activation, they release three different categories of molecules [231]. These molecules include (1) mediators that are constitutively stored in granules: histamine, serotonin, tryptase and heparin; (2) mediators that are synthesized in response to stimuli: leukotriene B4 (LTB4), prostaglandin (PG)D2 and lipid mediators; (3) cytokines and growth factors: TNF-α, IL-1β, IL-5, IL-8, granulocyte-macrophage colony-stimulating factor (GM-CSF), IL-10, VEGF and TGF-β [232,233,234,235]. Even with the current evidence on the significant role of mast cells in tissue repair, these cells showed heterogeneity in their functions depending on the tissue they reside in [236]. In addition, their phenotype changes according to the surrounding microenvironment [237]. Therefore, diverse mast cell subtypes with distinctive functions are yet to be characterized.

3.5. T Cells

Particular subtypes of T cells were shown to play a crucial role in tissue healing. For instance, dendritic epidermal T cells (DETCs) enhance re-epithelialization and granulation tissue formation via releasing various cytokines and growth factors [251,252]; consequently, mice with depleted DETCs experienced delayed wound closure and impaired ECM deposition [252,253,254]. Mechanistically, DETCs are activated by ligands (e.g., Semaphorin 4D and SKINTs) that are secreted mainly by keratinocytes upon injury [254,255]. Interestingly, stimulated DETCs express cytokines that are involved primarily in promoting keratinocyte proliferation, such as insulin growth factor-1 (IGF-1) and keratinocyte growth factors (KGF) [252,256]. Yet, the role of DETCs in several other tissue repair events is still undetermined.
Another type of T cell that is engaged directly in tissue repair is Treg. Several subsets of Treg were detected in various peripheral non-lymphoid organs [257]. The core functions of these cells are to negate detrimental inflammation and maintain tissue homeostasis [258,259]. Treg was found to reside in human and murine healthy skin, indicating their potential role in immunosurveillance and reparative responses in cases of injuries [260,261,262]. Treg predominantly mediates immune suppressive activities during tissue repair, where they contribute to suppressing inflammation and the transition toward the proliferative phase [263,264]. The immune-suppressive actions of Treg were attributed to their inhibition of IFN-γ+ T effector cells and proinflammatory Ly6C+ monocytes [265]. Furthermore, Treg directly regulates the polarization of anti-inflammatory/reparative M2 macrophages via expressing IL-13 and IL-4 [266] in addition to enhancing efferocytosis [267].

4. Suppression of Inflammation

Following eradication of infection and exclusion of cellular debris, a transition towards an anti-inflammatory program is essential for activating reparative pathways that restore the structure and function of damaged tissue. The process is achieved through a variety of suppressive signals prompting a reduction in proinflammatory mediators and infiltrating leukocytes in addition to upregulation of proresolution molecules, including IL-10 and TGF-β [268]. Several pathways were reported to regulate the resolution of inflammation, and they rely primarily on effectively eliminating microbes [269]. Defects in pathogen clearance necessitate a continuation of a proinflammatory reaction that ultimately results in delayed healing. The process further involves a cross-talk and interplay between immune and non-immune cells at the injury site.
Pathways regulating the control of inflammation can be categorized into cell- and cytokine-mediated mechanisms. The launch of resolution of acute inflammatory responses is likely to be timely mapped with the fading of PMN from the injury site [26]. Downregulation of PMN can be achieved through two main mechanisms: (1) apoptosis followed by efferocytosis and (2) reverse migration. We previously highlighted the role of neutrophils in resolving inflammation via their efferocytosis by macrophages [270]. The process is accomplished by binding cellular communication network factor 1 (CCN1), present on phosphatidylserine of apoptotic PMN, to integrins of macrophages [271]. Engulfing apoptotic neutrophils is critical to avoiding their secondary necrosis, which leads to a substantial release of detrimental proinflammatory cytokines and ROS [270], and subsequently transforming macrophages into an anti-inflammatory phenotype [135] (Figure 1). On the other hand, recent data indicates a retrograde migration of neutrophils back into circulation as a pathway of PMN resolution [150,272] (Figure 1). This was shown in various models of mice, zebrafish, and in vitro human neutrophils. Notably, prolonged inflammatory conditions have been commonly associated with extensive and prolonged neutrophil infiltration, resulting in chronic wounds [273]. Additionally, M2 macrophages control inflammation by secreting various anti-inflammatory mediators (Table 2). Other cell types at the injury site were reported to express proresolution cytokines such as IL-4, IL10, IL-13, IL-35 and TGF-β [274,275]. These cytokines suppress inflammation by inhibiting the synthesis of proinflammatory cytokine and chemokine [269,276]. Furthermore, they reduce cellular infiltration by repressing the expression of adhesion molecule and diminishing chemokine-mediated leukocyte recruitment [269,277].

5. Dysregulation of Inflammatory Responses and Its Outcome

Tight regulation of acute inflammation is critical for normal tissue repair. We previously emphasized the significance of cellular and cytokine effectors in the induction of acute inflammation. Dysregulation of these inflammatory responses eventually disrupts the healing process due to failure to transition to the proliferative phase (Figure 2). This involves either impairment or exaggeration of inflammatory mechanisms such as leukocyte recruitment and production of proinflammatory mediators. Therefore, a balance has to be maintained and resolution must be achieved in a timely manner to avoid extravagant inflammatory responses and their associated collateral tissue damage.
Non-healing injuries are commonly accompanied by persistent inflammation. Mechanistically, several factors were described to explain this phenomenon. For example, deregulated proteolytic activities (e.g., overproduction of proteoses) in uncontrolled inflammatory reactions can devastate protective repair mechanisms, including cleaving growth factors [278,279,280]. Activity and expression of various MMPs were substantially upregulated in chronic wounds [281]. Another factor is extensive and persistent neutrophil infiltration. Compromised resolution of PMN at the injury site is escorted by detrimental levels of ROS and proteases inducing damage to cell membranes, ECM and crucial tissue repair mediators such as TGF-β and PDGF [282,283]. Likewise, macrophages in chronic wounds are associated with reduced levels of tissue inhibitors of MMPs (TIMPs), thus augmenting ECM degradation and delaying healing [284]. Macrophages also tend to present a dysregulated expression of inflammatory mediators and growth factors in non-healing injuries [9]. This is further complicated by an imbalance in M1/M2 phenotype where alternatively activated macrophages are significantly diminished [285]. Conversely, Keratinocytes show impaired migration and proliferation abilities in chronic injuries [286,287,288]. Likewise, fibroblasts suffer the loss of their proliferative potentials due to being less responsive to growth factors [289].
Based on the previously mentioned observations, it is generally agreed that a proinflammatory cycle must be broken in order for non-healing injuries to heal properly. Therefore, to develop therapeutic measures that shield restored tissue from the detriment induced by persistent inflammatory microenvironments, it is crucial to unravel and fully characterize pro- and anti-inflammatory pathways in tissue repair [290].

6. Current Tissue Engineering Strategies Managing Chronic Injuries

During the past two decades, various therapeutic strategies employing tissue engineering technologies were introduced in the field of tissue repair as reviewed by Yu et al. [291]. These applications advanced progressively to include auto/allografts, engineered skin grafts, cell-based therapy, cytokine/growth factor delivery and modern multifunctional biomaterial-based dressings such as carbon nanomaterials, hydrogel, fibrous scaffold, sponge, acellular dermal matrix and foam. Herein, we focused on strategies modulating the inflammatory milieu and immune system behaviour through molecule analogs and signaling ligands released locally by engineered constructs. In chronic wounds, M1 (proinflammatory) phenotype fails to polarize to M2 (anti-inflammatory), leading to the persistent release of TNF-α and IL-1β, thus maintaining a state of chronic inflammation [202]. One of the currently utilized strategies is to activate polarization of these cells. For example, monocyte and macrophages, when exposed to sphingosine-1-phosphate in vitro, they preferentially retain anti-inflammatory phenotypes [292,293]. A polyvinyl alcohol sponge implant injected with sphingosine-1-phosphate and ciclopirox olamine (antifungal) demonstrated proangiogenic properties in diabetic rats [294]. Polymers such as poly lactic-co-glycolic acid with the capacity to control release of fingolimod enhanced recruitment of anti-inflammatory M2 macrophages and monocytes through stromal cell-derived factor-1 alpha-mediated chemotaxis [295]. Likewise, the properties of Keratin biomaterials and dextran isocyanatoethyl methacrylate ethylamine hydrogel were found to influence immune cell behaviour and responses, including M2 polarization [296,297]. These attributes were combined with promotion of overall wound healing and hair follicle formation.
Others considered using natural remedies that directly inhibit the chronic inflammatory cycle. For instance, a cellulose nanocrystal film releasing curcumin, a polyphenol with anti-inflammatory and antimicrobial effects, was found to enhance bacterial clearance and overall wound healing in diabetic rats [298]. Similarly, fabricated chitosan-sodium hyaluronate-resveratrol sponges sustaining resveratrol release into wounds promoted neutrophil resolution, granulation tissue formation, re-epithelialization and angiogenesis [299]. Directly inhibiting inflammatory signals locally could be a potentially effective strategy in managing chronic wounds. Nanoparticles associated with siRNA were reported to significantly reduce TNF-α and MCP-1 production by macrophages and fibroblasts, respectively [300]. These strategies avoid systemic anti-TNF-α therapeutic applications that risk the development of global immunosuppression.
Surface chemistry and topographical patterning, among other biomaterial modulations, can alter leukocyte microenvironments and their phenotypes [301]. The engineering of hydrogel with various peptide motifs creates immunomodulatory scaffolds that influence inflammatory responses at the injury site. For instance, fabricated hydrogel containing bioactive peptides (e.g., substance P) was reported to recruit mesenchymal stem cells (MSCs) that modulate inflammation intensity and promote Treg generation [302]. Future research aims at characterizing modulatory effects of several materials and biochemical factors may open the door toward robust therapeutic applications for chronic injuries.

7. Concluding Remarks

Despite the critical role of acute inflammation in tissue healing, a delicate regulation of its complex interacting network of diverse leukocyte subsets along with various pro- and anti-inflammatory mediators has to be maintained for efficient restoration of tissue homeostasis devoid of extensive collateral tissue injury. Importantly, a shift between the inflammatory and proliferative phases is crucial to preclude the unnecessary persistence of inflammation at the injury site. PMN apoptosis and their efferocytosis by macrophages are among the key signals inducing this transition. Pathologic and chronic non-healing injuries are mainly caused by the dysregulation of critical cellular and cytokine effectors during an acute inflammatory reaction, which lessens the effectiveness of the healing process. Therefore, it is necessary to uncover more about the molecular and cell-mediated mechanisms that regulate the inflammatory phase of tissue repair. This will provide insights that may open the door to novel therapeutic applications and strategies achieved via fine-tuning or enhancing these inflammatory pathways.

Author Contributions

Conceptualization, D.R.B. and A.M.S.; writing—original draft preparation, A.M.S.; writing—review and editing, D.R.B. and A.M.S.; supervision and project administration, D.R.B.; funding acquisition, D.R.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a Natural Sciences and Engineering Council of Canada grant to Daniel Barreda (RGPIN-2018-05768).

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Galis, F.; Wagner, G.P.; Jockusch, E.L. Why Is Limb Regeneration Possible in Amphibians but Not in Reptiles, Birds, and Mammals? Evol. Dev. 2003, 5, 208–220. [Google Scholar] [CrossRef] [PubMed]
  2. Frangogiannis, N.G. The Inflammatory Response in Tissue Repair. In Inflammation; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2017; pp. 1517–1538. ISBN 978-3-527-69215-6. [Google Scholar]
  3. Soliman, A.M.; Yoon, T.; Wang, J.; Stafford, J.L.; Barreda, D.R. Isolation of Skin Leukocytes Uncovers Phagocyte Inflammatory Responses During Induction and Resolution of Cutaneous Inflammation in Fish. Front. Immunol. 2021, 12, 725063. [Google Scholar] [CrossRef]
  4. Krafts, K.P. Tissue Repair. Organogenesis 2010, 6, 225–233. [Google Scholar] [CrossRef] [Green Version]
  5. Clark, R.A.F. Fibrin Is a Many Splendored Thing. J. Investig. Dermatol. 2003, 121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Wilgus, T.A.; Roy, S.; McDaniel, J.C. Neutrophils and Wound Repair: Positive Actions and Negative Reactions. Adv. Wound Care 2013, 2, 379–388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Park, J.E.; Barbul, A. Understanding the Role of Immune Regulation in Wound Healing. Am. J. Surg. 2004, 187, 11S–16S. [Google Scholar] [CrossRef] [PubMed]
  8. Brem, H.; Tomic-Canic, M. Cellular and Molecular Basis of Wound Healing in Diabetes. J. Clin. Investig. 2007, 117, 1219–1222. [Google Scholar] [CrossRef] [Green Version]
  9. Wynn, T.A.; Vannella, K.M. Macrophages in Tissue Repair, Regeneration, and Fibrosis. Immunity 2016, 44, 450–462. [Google Scholar] [CrossRef] [Green Version]
  10. Gurtner, G.C.; Werner, S.; Barrandon, Y.; Longaker, M.T. Wound Repair and Regeneration. Nature 2008, 453, 314–321. [Google Scholar] [CrossRef]
  11. Tonnesen, M.G.; Feng, X.; Clark, R.A.F. Angiogenesis in Wound Healing. J. Investig. Dermatol. Symp. Proc. 2000, 5, 40–46. [Google Scholar] [CrossRef]
  12. Gerhardt, H.; Golding, M.; Fruttiger, M.; Ruhrberg, C.; Lundkvist, A.; Abramsson, A.; Jeltsch, M.; Mitchell, C.; Alitalo, K.; Shima, D.; et al. VEGF Guides Angiogenic Sprouting Utilizing Endothelial Tip Cell Filopodia. J. Cell Biol. 2003, 161, 1163–1177. [Google Scholar] [CrossRef] [PubMed]
  13. Kaltalioglu, K.; Coskun-Cevher, S. A Bioactive Molecule in a Complex Wound Healing Process: Platelet-Derived Growth Factor. Int. J. Dermatol. 2015, 54, 972–977. [Google Scholar] [CrossRef]
  14. Xu, J.; Lamouille, S.; Derynck, R. TGF-β-Induced Epithelial to Mesenchymal Transition. Cell Res. 2009, 19, 156–172. [Google Scholar] [CrossRef] [PubMed]
  15. Werner, S.; Krieg, T.; Smola, H. Keratinocyte-Fibroblast Interactions in Wound Healing. J. Investig. Dermatol. 2007, 127, 998–1008. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Li, J.; Chen, J.; Kirsner, R. Pathophysiology of Acute Wound Healing. Clin. Dermatol. 2007, 25, 9–18. [Google Scholar] [CrossRef]
  17. Xue, M.; Jackson, C.J. Extracellular Matrix Reorganization During Wound Healing and Its Impact on Abnormal Scarring. Adv. Wound Care 2015, 4, 119–136. [Google Scholar] [CrossRef] [Green Version]
  18. Wilkinson, H.N.; Hardman, M.J. Wound Healing: Cellular Mechanisms and Pathological Outcomes. Open Biol. 2020, 10, 200223. [Google Scholar] [CrossRef]
  19. Singhal, P.K.; Sassi, S.; Lan, L.; Au, P.; Halvorsen, S.C.; Fukumura, D.; Jain, R.K.; Seed, B. Mouse Embryonic Fibroblasts Exhibit Extensive Developmental and Phenotypic Diversity. Proc. Natl. Acad. Sci. USA 2016, 113, 122–127. [Google Scholar] [CrossRef] [Green Version]
  20. Driskell, R.R.; Lichtenberger, B.M.; Hoste, E.; Kretzschmar, K.; Simons, B.D.; Charalambous, M.; Ferron, S.R.; Herault, Y.; Pavlovic, G.; Ferguson-Smith, A.C.; et al. Distinct Fibroblast Lineages Determine Dermal Architecture in Skin Development and Repair. Nature 2013, 504, 277–281. [Google Scholar] [CrossRef] [Green Version]
  21. Fries, K.M.; Blieden, T.; Looney, R.J.; Sempowski, G.D.; Silvera, M.R.; Willis, R.A.; Phipps, R.P. Evidence of Fibroblast Heterogeneity and the Role of Fibroblast Subpopulations in Fibrosis. Clin. Immunol. Immunopathol. 1994, 72, 283–292. [Google Scholar] [CrossRef]
  22. Donaldson, D.J.; Mahan, J.T. Fibrinogen and Fibronectin as Substrates for Epidermal Cell Migration during Wound Closure. J. Cell Sci. 1983, 62, 117–127. [Google Scholar] [CrossRef] [PubMed]
  23. Baum, C.L.; Arpey, C.J. Normal Cutaneous Wound Healing: Clinical Correlation with Cellular and Molecular Events. Dermatol. Surg. 2005, 31, 674–686. [Google Scholar] [CrossRef] [PubMed]
  24. Rousselle, P.; Braye, F.; Dayan, G. Re-Epithelialization of Adult Skin Wounds: Cellular Mechanisms and Therapeutic Strategies. Adv. Drug Deliv. Rev. 2019, 146, 344–365. [Google Scholar] [CrossRef] [PubMed]
  25. Santoro, M.M.; Gaudino, G. Cellular and Molecular Facets of Keratinocyte Reepithelization during Wound Healing. Exp. Cell Res. 2005, 304, 274–286. [Google Scholar] [CrossRef] [PubMed]
  26. Rodrigues, M.; Kosaric, N.; Bonham, C.A.; Gurtner, G.C. Wound Healing: A Cellular Perspective. Physiol. Rev. 2019, 99, 665–706. [Google Scholar] [CrossRef]
  27. Holzer-Geissler, J.C.J.; Schwingenschuh, S.; Zacharias, M.; Einsiedler, J.; Kainz, S.; Reisenegger, P.; Holecek, C.; Hofmann, E.; Wolff-Winiski, B.; Fahrngruber, H.; et al. The Impact of Prolonged Inflammation on Wound Healing. Biomedicines 2022, 10, 856. [Google Scholar] [CrossRef]
  28. Zhao, R.; Liang, H.; Clarke, E.; Jackson, C.; Xue, M. Inflammation in Chronic Wounds. Int. J. Mol. Sci. 2016, 17, 2085. [Google Scholar] [CrossRef]
  29. Jung, K.; Covington, S.; Sen, C.K.; Januszyk, M.; Kirsner, R.S.; Gurtner, G.C.; Shah, N.H. Rapid Identification of Slow Healing Wounds. Wound Repair Regen. 2016, 24, 181–188. [Google Scholar] [CrossRef] [Green Version]
  30. Roh, J.S.; Sohn, D.H. Damage-Associated Molecular Patterns in Inflammatory Diseases. Immune Netw. 2018, 18, e27. [Google Scholar] [CrossRef]
  31. Mogensen, T.H. Pathogen Recognition and Inflammatory Signaling in Innate Immune Defenses. Clin. Microbiol. Rev. 2009, 22, 240–273. [Google Scholar] [CrossRef]
  32. Abdallah, F.; Mijouin, L.; Pichon, C. Skin Immune Landscape: Inside and Outside the Organism. Mediat. Inflamm. 2017, 2017, 5095293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Pradeu, T.; Cooper, E.L. The Danger Theory: 20 Years Later. Front. Immunol. 2012, 3, 287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Kono, H.; Rock, K.L. How Dying Cells Alert the Immune System to Danger. Nat. Rev. Immunol. 2008, 8, 279–289. [Google Scholar] [CrossRef] [PubMed]
  35. Ishii, K.J.; Suzuki, K.; Coban, C.; Takeshita, F.; Itoh, Y.; Matoba, H.; Kohn, L.D.; Klinman, D.M. Genomic DNA Released by Dying Cells Induces the Maturation of APCs. J. Immunol. 2001, 167, 2602–2607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Basu, S.; Binder, R.J.; Suto, R.; Anderson, K.M.; Srivastava, P.K. Necrotic but Not Apoptotic Cell Death Releases Heat Shock Proteins, Which Deliver a Partial Maturation Signal to Dendritic Cells and Activate the NF-Kappa B Pathway. Int. Immunol. 2000, 12, 1539–1546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Shi, Y.; Evans, J.E.; Rock, K.L. Molecular Identification of a Danger Signal That Alerts the Immune System to Dying Cells. Nature 2003, 425, 516–521. [Google Scholar] [CrossRef] [PubMed]
  38. Wenceslau, C.F.; McCarthy, C.G.; Szasz, T.; Spitler, K.; Goulopoulou, S.; Webb, R.C. Working Group on DAMPs in Cardiovascular Disease Mitochondrial Damage-Associated Molecular Patterns and Vascular Function. Eur. Heart J. 2014, 35, 1172–1177. [Google Scholar] [CrossRef] [Green Version]
  39. Weathington, N.M.; van Houwelingen, A.H.; Noerager, B.D.; Jackson, P.L.; Kraneveld, A.D.; Galin, F.S.; Folkerts, G.; Nijkamp, F.P.; Blalock, J.E. A Novel Peptide CXCR Ligand Derived from Extracellular Matrix Degradation during Airway Inflammation. Nat. Med. 2006, 12, 317–323. [Google Scholar] [CrossRef]
  40. Taylor, K.R.; Trowbridge, J.M.; Rudisill, J.A.; Termeer, C.C.; Simon, J.C.; Gallo, R.L. Hyaluronan Fragments Stimulate Endothelial Recognition of Injury through TLR4. J. Biol. Chem. 2004, 279, 17079–17084. [Google Scholar] [CrossRef] [Green Version]
  41. Wrenshall, L.E.; Cerra, F.B.; Carlson, A.; Bach, F.H.; Platt, J.L. Regulation of Murine Splenocyte Responses by Heparan Sulfate. J. Immunol. 1991, 147, 455–459. [Google Scholar]
  42. Van der Vliet, A.; Janssen-Heininger, Y.M.W. Hydrogen Peroxide as a Damage Signal in Tissue Injury and Inflammation: Murderer, Mediator, or Messenger? J. Cell. Biochem. 2014, 115, 427–435. [Google Scholar] [CrossRef] [PubMed]
  43. Beutler, B. Tlr4: Central Component of the Sole Mammalian LPS Sensor. Curr. Opin. Immunol. 2000, 12, 20–26. [Google Scholar] [CrossRef] [PubMed]
  44. Hashimoto, M.; Tawaratsumida, K.; Kariya, H.; Aoyama, K.; Tamura, T.; Suda, Y. Lipoprotein Is a Predominant Toll-like Receptor 2 Ligand in Staphylococcus Aureus Cell Wall Components. Int. Immunol. 2006, 18, 355–362. [Google Scholar] [CrossRef]
  45. Levitz, S.M. Innate Recognition of Fungal Cell Walls. PLoS Pathog. 2010, 6, e1000758. [Google Scholar] [CrossRef] [Green Version]
  46. Paludan, S.R.; Bowie, A.G. Immune Sensing of DNA. Immunity 2013, 38, 870–880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Deane, J.A.; Bolland, S. Nucleic Acid-Sensing TLRs as Modifiers of Autoimmunity. J. Immunol. 2006, 177, 6573–6578. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Li, D.; Wu, M. Pattern Recognition Receptors in Health and Diseases. Signal Transduct. Target. Ther. 2021, 6, e17023. [Google Scholar] [CrossRef]
  49. Kumar, H.; Kawai, T.; Akira, S. Toll-like Receptors and Innate Immunity. Biochem. Biophys. Res. Commun. 2009, 388, 621–625. [Google Scholar] [CrossRef]
  50. Tsirogianni, A.K.; Moutsopoulos, N.M.; Moutsopoulos, H.M. Wound Healing: Immunological Aspects. Injury 2006, 37, S5–S12. [Google Scholar] [CrossRef]
  51. Zhang, F.X.; Kirschning, C.J.; Mancinelli, R.; Xu, X.-P.; Jin, Y.; Faure, E.; Mantovani, A.; Rothe, M.; Muzio, M.; Arditi, M. Bacterial Lipopolysaccharide Activates Nuclear Factor-ΚB through Interleukin-1 Signaling Mediators in Cultured Human Dermal Endothelial Cells and Mononuclear Phagocytes. J. Biol. Chem. 1999, 274, 7611–7614. [Google Scholar] [CrossRef] [Green Version]
  52. Kaisho, T.; Akira, S. Toll-like Receptor Function and Signaling. J. Allergy Clin. Immunol. 2006, 117, 979–987. [Google Scholar] [CrossRef] [PubMed]
  53. Robert, C.; Kupper, T.S. Inflammatory Skin Diseases, T Cells, and Immune Surveillance. N. Engl. J. Med. 1999, 341, 1817–1828. [Google Scholar] [CrossRef] [PubMed]
  54. Lansdown, A.B.G. Calcium: A Potential Central Regulator in Wound Healing in the Skin. Wound Repair Regen. 2002, 10, 271–285. [Google Scholar] [CrossRef] [PubMed]
  55. Dinarello, C.A. Historical Review of Cytokines. Eur. J. Immunol. 2007, 37, S34–S45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Barrientos, S.; Stojadinovic, O.; Golinko, M.S.; Brem, H.; Tomic-Canic, M. Growth Factors and Cytokines in Wound Healing. Wound Repair Regen. 2008, 16, 585–601. [Google Scholar] [CrossRef]
  57. Gethin, G. Understanding the Inflammatory Process in Wound Healing. Br. J. Community Nurs. 2012, 17, S17–S22. [Google Scholar] [CrossRef]
  58. Kishimoto, T. The Biology of Interleukin-6. Blood 1989, 74, 1–10. [Google Scholar] [CrossRef] [Green Version]
  59. Lin, Z.-Q.; Kondo, T.; Ishida, Y.; Takayasu, T.; Mukaida, N. Essential Involvement of IL-6 in the Skin Wound-Healing Process as Evidenced by Delayed Wound Healing in IL-6-Deficient Mice. J. Leukoc. Biol. 2003, 73, 713–721. [Google Scholar] [CrossRef] [Green Version]
  60. Weller, K.; Foitzik, K.; Paus, R.; Syska, W.; Maurer, M. Mast Cells Are Required for Normal Healing of Skin Wounds in Mice. FASEB J. 2006, 20, 2366–2368. [Google Scholar] [CrossRef]
  61. Garbuzenko, E.; Nagler, A.; Pickholtz, D.; Gillery, P.; Reich, R.; Maquart, F.-X.; Levi-Schaffer, F. Human Mast Cells Stimulate Fibroblast Proliferation, Collagen Synthesis and Lattice Contraction: A Direct Role for Mast Cells in Skin Fibrosis. Clin. Exp. Allergy 2002, 32, 237–246. [Google Scholar] [CrossRef]
  62. Blair, P.; Flaumenhaft, R. Platelet Alpha-Granules: Basic Biology and Clinical Correlates. Blood Rev. 2009, 23, 177–189. [Google Scholar] [CrossRef] [PubMed]
  63. Zhang, S.; Liu, Y.; Zhang, X.; Zhu, D.; Qi, X.; Cao, X.; Fang, Y.; Che, Y.; Han, Z.-C.; He, Z.-X.; et al. Prostaglandin E2 Hydrogel Improves Cutaneous Wound Healing via M2 Macrophages Polarization. Theranostics 2018, 8, 5348–5361. [Google Scholar] [CrossRef] [PubMed]
  64. Takeuchi, K.; Tanigami, M.; Amagase, K.; Ochi, A.; Okuda, S.; Hatazawa, R. Endogenous Prostaglandin E2 Accelerates Healing of Indomethacin-Induced Small Intestinal Lesions through Upregulation of Vascular Endothelial Growth Factor Expression by Activation of EP4 Receptors. J. Gastroenterol. Hepatol. 2010, 25 (Suppl. S1), S67–S74. [Google Scholar] [CrossRef]
  65. Zhang, Q.-Z.; Su, W.-R.; Shi, S.-H.; Wilder-Smith, P.; Xiang, A.P.; Wong, A.; Nguyen, A.L.; Kwon, C.W.; Le, A.D. Human Gingiva-Derived Mesenchymal Stem Cells Elicit Polarization of M2 Macrophages and Enhance Cutaneous Wound Healing. Stem Cells 2010, 28, 1856–1868. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Crunkhorn, S. Regenerative Medicine: Inhibiting Prostaglandin Breakdown Triggers Tissue Regeneration. Nat. Rev. Drug Discov. 2015, 14, 526. [Google Scholar] [CrossRef] [PubMed]
  67. Ogletree, M.L. Overview of Physiological and Pathophysiological Effects of Thromboxane A2. Fed. Proc. 1987, 46, 133–138. [Google Scholar]
  68. Pierre, S.; Linke, B.; Suo, J.; Tarighi, N.; Del Turco, D.; Thomas, D.; Ferreiros, N.; Stegner, D.; Frölich, S.; Sisignano, M.; et al. GPVI and Thromboxane Receptor on Platelets Promote Proinflammatory Macrophage Phenotypes during Cutaneous Inflammation. J. Investig. Dermatol. 2017, 137, 686–695. [Google Scholar] [CrossRef] [Green Version]
  69. Daniel, T.O.; Liu, H.; Morrow, J.D.; Crews, B.C.; Marnett, L.J. Thromboxane A2 Is a Mediator of Cyclooxygenase-2-Dependent Endothelial Migration and Angiogenesis. Cancer Res. 1999, 59, 4574–4577. [Google Scholar]
  70. Yokomizo, T.; Izumi, T.; Chang, K.; Takuwa, Y.; Shimizu, T. A G-Protein-Coupled Receptor for Leukotriene B4 That Mediates Chemotaxis. Nature 1997, 387, 620–624. [Google Scholar] [CrossRef]
  71. Brandt, S.L.; Wang, S.; Dejani, N.N.; Klopfenstein, N.; Winfree, S.; Filgueiras, L.; McCarthy, B.P.; Territo, P.R.; Serezani, C.H. Excessive Localized Leukotriene B4 Levels Dictate Poor Skin Host Defense in Diabetic Mice. JCI Insight 2018, 3, e120220. [Google Scholar] [CrossRef] [Green Version]
  72. Raziyeva, K.; Kim, Y.; Zharkinbekov, Z.; Kassymbek, K.; Jimi, S.; Saparov, A. Immunology of Acute and Chronic Wound Healing. Biomolecules 2021, 11, 700. [Google Scholar] [CrossRef] [PubMed]
  73. Singer, A.J.; Clark, R.A. Cutaneous Wound Healing. N. Engl. J. Med. 1999, 341, 738–746. [Google Scholar] [CrossRef] [PubMed]
  74. Nosenko, M.A.; Ambaryan, S.G.; Drutskaya, M.S. Proinflammatory Cytokines and Skin Wound Healing in Mice. Mol. Biol. 2019, 53, 653–664. [Google Scholar] [CrossRef]
  75. Frank, J.; Born, K.; Barker, J.H.; Marzi, I. In Vivo Effect of Tumor NecrosisFactor Alpha on Wound Angiogenesis AndEpithelialization. Eur. J. Trauma 2003, 29, 208–219. [Google Scholar] [CrossRef]
  76. Shinozaki, M.; Okada, Y.; Kitano, A.; Ikeda, K.; Saika, S.; Shinozaki, M. Impaired Cutaneous Wound Healing with Excess Granulation Tissue Formation in TNFalpha-Null Mice. Arch Dermatol. Res. 2009, 301, 531–537. [Google Scholar] [CrossRef]
  77. Dudeck, J.; Froebel, J.; Kotrba, J.; Lehmann, C.H.K.; Dudziak, D.; Speier, S.; Nedospasov, S.A.; Schraven, B.; Dudeck, A. Engulfment of Mast Cell Secretory Granules on Skin Inflammation Boosts Dendritic Cell Migration and Priming Efficiency. J. Allergy Clin. Immunol. 2019, 143, 1849–1864.e4. [Google Scholar] [CrossRef]
  78. Qing, C. The Molecular Biology in Wound Healing & Non-Healing Wound. Chin. J. Traumatol. 2017, 20, 189–193. [Google Scholar] [CrossRef]
  79. Xue, X.; Falcon, D.M. The Role of Immune Cells and Cytokines in Intestinal Wound Healing. Int. J. Mol. Sci. 2019, 20, 6097. [Google Scholar] [CrossRef] [Green Version]
  80. Kawasaki, Y.; Zhang, L.; Cheng, J.-K.; Ji, R.-R. Cytokine Mechanisms of Central Sensitization: Distinct and Overlapping Role of Interleukin-1beta, Interleukin-6, and Tumor Necrosis Factor-Alpha in Regulating Synaptic and Neuronal Activity in the Superficial Spinal Cord. J. Neurosci. 2008, 28, 5189–5194. [Google Scholar] [CrossRef] [Green Version]
  81. Werner, S.; Peters, K.G.; Longaker, M.T.; Fuller-Pace, F.; Banda, M.J.; Williams, L.T. Large Induction of Keratinocyte Growth Factor Expression in the Dermis during Wound Healing. Proc. Natl. Acad. Sci. USA 1992, 89, 6896–6900. [Google Scholar] [CrossRef] [Green Version]
  82. Tang, A.; Gilchrest, B.A. Regulation of Keratinocyte Growth Factor Gene Expression in Human Skin Fibroblasts. J. Dermatol. Sci. 1996, 11, 41–50. [Google Scholar] [CrossRef] [PubMed]
  83. Lee, P.; Gund, R.; Dutta, A.; Pincha, N.; Rana, I.; Ghosh, S.; Witherden, D.; Kandyba, E.; MacLeod, A.; Kobielak, K.; et al. Stimulation of Hair Follicle Stem Cell Proliferation through an IL-1 Dependent Activation of ΓδT-Cells. eLife 2017, 6, e28875. [Google Scholar] [CrossRef] [PubMed]
  84. Turner, N.A. Effects of Interleukin-1 on Cardiac Fibroblast Function: Relevance to Post-Myocardial Infarction Remodelling. Vasc. Pharmacol. 2014, 60, 1–7. [Google Scholar] [CrossRef] [PubMed]
  85. Mitchell, M.D.; Laird, R.E.; Brown, R.D.; Long, C.S. IL-1beta Stimulates Rat Cardiac Fibroblast Migration via MAP Kinase Pathways. Am. J. Physiol. Heart Circ. Physiol. 2007, 292, H1139–H1147. [Google Scholar] [CrossRef]
  86. De Oliveira, S.; Reyes-Aldasoro, C.C.; Candel, S.; Renshaw, S.A.; Mulero, V.; Calado, Â. Cxcl8 (Interleukin-8) Mediates Neutrophil Recruitment and Behavior in the Zebrafish Inflammatory Response. J. Immunol. 2013, 190, 4349–4359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Das, S.T.; Rajagopalan, L.; Guerrero-Plata, A.; Sai, J.; Richmond, A.; Garofalo, R.P.; Rajarathnam, K. Monomeric and Dimeric CXCL8 Are Both Essential for In Vivo Neutrophil Recruitment. PLoS ONE 2010, 5, e11754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Paccaud, J.P.; Schifferli, J.A.; Baggiolini, M. NAP-1/IL-8 Induces up-Regulation of CR1 Receptors in Human Neutrophil Leukocytes. Biochem. Biophys. Res. Commun. 1990, 166, 187–192. [Google Scholar] [CrossRef] [PubMed]
  89. Cowland, J.B.; Borregaard, N. Granulopoiesis and Granules of Human Neutrophils. Immunol. Rev. 2016, 273, 11–28. [Google Scholar] [CrossRef]
  90. Rabe, B.; Chalaris, A.; May, U.; Waetzig, G.H.; Seegert, D.; Williams, A.S.; Jones, S.A.; Rose-John, S.; Scheller, J. Transgenic Blockade of Interleukin 6 Transsignaling Abrogates Inflammation. Blood 2008, 111, 1021–1028. [Google Scholar] [CrossRef]
  91. Rose-John, S. IL-6 Trans-Signaling via the Soluble IL-6 Receptor: Importance for the pro-Inflammatory Activities of IL-6. Int. J. Biol. Sci. 2012, 8, 1237–1247. [Google Scholar] [CrossRef] [Green Version]
  92. Yang, R.; Masters, A.R.; Fortner, K.A.; Champagne, D.P.; Yanguas-Casás, N.; Silberger, D.J.; Weaver, C.T.; Haynes, L.; Rincon, M. IL-6 Promotes the Differentiation of a Subset of Naive CD8+ T Cells into IL-21-Producing B Helper CD8+ T Cells. J. Exp. Med. 2016, 213, 2281–2291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Bosurgi, L.; Cao, Y.G.; Cabeza-Cabrerizo, M.; Tucci, A.; Hughes, L.D.; Kong, Y.; Weinstein, J.S.; Licona-Limon, P.; Schmid, E.T.; Pelorosso, F.; et al. Macrophage Function in Tissue Repair and Remodeling Requires IL-4 or IL-13 with Apoptotic Cells. Science 2017, 356, 1072–1076. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Luckett-Chastain, L.R.; Gallucci, R.M. Interleukin (IL)-6 Modulates Transforming Growth Factor-β Expression in Skin and Dermal Fibroblasts from IL-6-Deficient Mice. Br. J. Dermatol. 2009, 161, 237–248. [Google Scholar] [CrossRef] [PubMed]
  95. McFarland-Mancini, M.M.; Funk, H.M.; Paluch, A.M.; Zhou, M.; Giridhar, P.V.; Mercer, C.A.; Kozma, S.C.; Drew, A.F. Differences in Wound Healing in Mice with Deficiency of IL-6 versus IL-6 Receptor. J. Immunol. 2010, 184, 7219–7228. [Google Scholar] [CrossRef] [Green Version]
  96. Luckett, L.R.; Gallucci, R.M. Interleukin-6 (IL-6) Modulates Migration and Matrix Metalloproteinase Function in Dermal Fibroblasts from IL-6KO Mice. Br. J. Dermatol. 2007, 156, 1163–1171. [Google Scholar] [CrossRef]
  97. Gallucci, R.M.; Sugawara, T.; Yucesoy, B.; Berryann, K.; Simeonova, P.P.; Matheson, J.M.; Luster, M.I. Interleukin-6 Treatment Augments Cutaneous Wound Healing in Immunosuppressed Mice. J. Interferon Cytokine Res. 2001, 21, 603–609. [Google Scholar] [CrossRef]
  98. Johnson, B.Z.; Stevenson, A.W.; Prêle, C.M.; Fear, M.W.; Wood, F.M. The Role of IL-6 in Skin Fibrosis and Cutaneous Wound Healing. Biomedicines 2020, 8, 101. [Google Scholar] [CrossRef]
  99. Brown, L.F.; Yeo, K.T.; Berse, B.; Yeo, T.K.; Senger, D.R.; Dvorak, H.F.; van de Water, L. Expression of Vascular Permeability Factor (Vascular Endothelial Growth Factor) by Epidermal Keratinocytes during Wound Healing. J. Exp. Med. 1992, 176, 1375–1379. [Google Scholar] [CrossRef] [Green Version]
  100. Schoenborn, J.R.; Wilson, C.B. Regulation of Interferon-γ During Innate and Adaptive Immune Responses. In Advances in Immunology; Academic Press: Cambridge, MA, USA, 2007; Volume 96, pp. 41–101. [Google Scholar]
  101. Miles, R.H.; Paxton, T.P.; Zacheis, D.; Dries, D.J.; Gamelli, R.L. Systemic Administration of Interferon-Gamma Impairs Wound Healing. J. Surg. Res. 1994, 56, 288–294. [Google Scholar] [CrossRef]
  102. Ishida, Y.; Kondo, T.; Takayasu, T.; Iwakura, Y.; Mukaida, N. The Essential Involvement of Cross-Talk between IFN-γ and TGF-β in the Skin Wound-Healing Process. J. Immunol. 2004, 172, 1848–1855. [Google Scholar] [CrossRef] [Green Version]
  103. King, A.; Balaji, S.; Le, L.D.; Crombleholme, T.M.; Keswani, S.G. Regenerative Wound Healing: The Role of Interleukin-10. Adv. Wound Care 2014, 3, 315–323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Saraiva, M.; Vieira, P.; O’Garra, A. Biology and Therapeutic Potential of Interleukin-10. J. Exp. Med. 2019, 217, e20190418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. King, A.; Balaji, S.; Le, L.D.; Marsh, E.; Crombleholme, T.M.; Keswani, S.G. Interleukin-10 Regulates Fetal Extracellular Matrix Hyaluronan Production. J. Pediatr. Surg. 2013, 48, 1211–1217. [Google Scholar] [CrossRef] [PubMed]
  106. Shi, J.-H.; Guan, H.; Shi, S.; Cai, W.-X.; Bai, X.-Z.; Hu, X.-L.; Fang, X.-B.; Liu, J.-Q.; Tao, K.; Zhu, X.-X.; et al. Protection against TGF-Β1-Induced Fibrosis Effects of IL-10 on Dermal Fibroblasts and Its Potential Therapeutics for the Reduction of Skin Scarring. Arch. Dermatol. Res. 2013, 305, 341–352. [Google Scholar] [CrossRef]
  107. Li, M.O.; Wan, Y.Y.; Sanjabi, S.; Robertson, A.-K.L.; Flavell, R.A. Transforming Growth Factor-Beta Regulation of Immune Responses. Annu. Rev. Immunol. 2006, 24, 99–146. [Google Scholar] [CrossRef]
  108. Reibman, J.; Meixler, S.; Lee, T.C.; Gold, L.I.; Cronstein, B.N.; Haines, K.A.; Kolasinski, S.L.; Weissmann, G. Transforming Growth Factor Beta 1, a Potent Chemoattractant for Human Neutrophils, Bypasses Classic Signal-Transduction Pathways. Proc. Natl. Acad. Sci. USA 1991, 88, 6805–6809. [Google Scholar] [CrossRef] [Green Version]
  109. White, L.A.; Mitchell, T.I.; Brinckerhoff, C.E. Transforming Growth Factor Beta Inhibitory Element in the Rabbit Matrix Metalloproteinase-1 (Collagenase-1) Gene Functions as a Repressor of Constitutive Transcription. Biochim. Biophys. Acta 2000, 1490, 259–268. [Google Scholar] [CrossRef]
  110. Evrard, S.M.; d’Audigier, C.; Mauge, L.; Israël-Biet, D.; Guerin, C.L.; Bieche, I.; Kovacic, J.C.; Fischer, A.-M.; Gaussem, P.; Smadja, D.M. The Profibrotic Cytokine Transforming Growth Factor-Β1 Increases Endothelial Progenitor Cell Angiogenic Properties. J. Thromb. Haemost. 2012, 10, 670–679. [Google Scholar] [CrossRef]
  111. Ramirez, H.; Patel, S.B.; Pastar, I. The Role of TGFβ Signaling in Wound Epithelialization. Adv. Wound Care 2014, 3, 482–491. [Google Scholar] [CrossRef] [Green Version]
  112. Hinz, B. Formation and Function of the Myofibroblast during Tissue Repair. J. Investig. Dermatol. 2007, 127, 526–537. [Google Scholar] [CrossRef]
  113. Gilbert, R.W.D.; Vickaryous, M.K.; Viloria-Petit, A.M. Signalling by Transforming Growth Factor Beta Isoforms in Wound Healing and Tissue Regeneration. J. Dev. Biol. 2016, 4, 21. [Google Scholar] [CrossRef] [Green Version]
  114. Martins-Green, M.; Petreaca, M.; Wang, L. Chemokines and Their Receptors Are Key Players in the Orchestra That Regulates Wound Healing. Adv. Wound Care 2013, 2, 327–347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Wang, J.; Knaut, H. Chemokine Signaling in Development and Disease. Development 2014, 141, 4199–4205. [Google Scholar] [CrossRef] [PubMed]
  116. Clark-Lewis, I.; Kim, K.S.; Rajarathnam, K.; Gong, J.H.; Dewald, B.; Moser, B.; Baggiolini, M.; Sykes, B.D. Structure-Activity Relationships of Chemokines. J. Leukoc. Biol. 1995, 57, 703–711. [Google Scholar] [CrossRef] [PubMed]
  117. Zlotnik, A.; Morales, J.; Hedrick, J.A. Recent Advances in Chemokines and Chemokine Receptors. Crit. Rev. Immunol. 1999, 19, 47. [Google Scholar] [CrossRef]
  118. Moser, B.; Dewald, B.; Barella, L.; Schumacher, C.; Baggiolini, M.; Clark-Lewis, I. Interleukin-8 Antagonists Generated by N-Terminal Modification. J. Biol. Chem. 1993, 268, 7125–7128. [Google Scholar] [CrossRef] [PubMed]
  119. Frangogiannis, N.G. Chemokines in Ischemia and Reperfusion. Thromb. Haemost. 2007, 97, 738–747. [Google Scholar] [CrossRef]
  120. Miller, M.D.; Krangel, M.S. The Human Cytokine I-309 Is a Monocyte Chemoattractant. Proc. Natl. Acad. Sci. USA 1992, 89, 2950–2954. [Google Scholar] [CrossRef] [Green Version]
  121. D’Ambrosio, D.; Iellem, A.; Bonecchi, R.; Mazzeo, D.; Sozzani, S.; Mantovani, A.; Sinigaglia, F. Cutting Edge: Selective Up-Regulation of Chemokine Receptors CCR4 and CCR8 upon Activation of Polarized Human Type 2 Th Cells. J. Immunol. 1998, 161, 5111–5115. [Google Scholar]
  122. Singh, U.P.; Singh, N.P.; Murphy, E.A.; Price, R.L.; Fayad, R.; Nagarkatti, M.; Nagarkatti, P.S. Chemokine and Cytokine Levels in Inflammatory Bowel Disease Patients. Cytokine 2016, 77, 44–49. [Google Scholar] [CrossRef] [Green Version]
  123. Jin, T.; Xu, X.; Hereld, D. Chemotaxis, Chemokine Receptors and Human Disease. Cytokine 2008, 44, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Gerard, C.; Rollins, B.J. Chemokines and Disease. Nat. Immunol. 2001, 2, 108–115. [Google Scholar] [CrossRef] [PubMed]
  125. Soliman, A.M.; Sim, R.H.; Das, S.; Mahakkanukrauh, P. Therapeutic Targeting of Inflammatory Pathways with Emphasis on NLRP3 Inflammasomes by Natural Products: A Novel Approach for the Treatment of Inflammatory Eye Diseases. Curr. Med. Chem. 2022, 29, 2891–2912. [Google Scholar] [CrossRef] [PubMed]
  126. Soliman, A.M.; Das, S.; Mahakkanukrauh, P. Inflammatory Molecular Mediators and Pathways Involved in Vascular Aging and Stroke: A Comprehensive Review. Curr. Med. Chem. 2021, 28, 5522–5542. [Google Scholar] [CrossRef]
  127. Proudfoot, A.E.I.; Handel, T.M.; Johnson, Z.; Lau, E.K.; LiWang, P.; Clark-Lewis, I.; Borlat, F.; Wells, T.N.C.; Kosco-Vilbois, M.H. Glycosaminoglycan Binding and Oligomerization Are Essential for the in Vivo Activity of Certain Chemokines. Proc. Natl. Acad. Sci. USA 2003, 100, 1885–1890. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Rollins, B.J. Chemokines. Blood 1997, 90, 909–928. [Google Scholar] [CrossRef]
  129. Olson, T.S.; Ley, K. Chemokines and Chemokine Receptors in Leukocyte Trafficking. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2002, 283, R7–R28. [Google Scholar] [CrossRef] [Green Version]
  130. Russo, R.C.; Garcia, C.C.; Teixeira, M.M.; Amaral, F.A. The CXCL8/IL-8 Chemokine Family and Its Receptors in Inflammatory Diseases. Expert Rev. Clin. Immunol. 2014, 10, 593–619. [Google Scholar] [CrossRef] [Green Version]
  131. Zaja-Milatovic, S.; Richmond, A. CXC Chemokines and Their Receptors: A Case for a Significant Biological Role in Cutaneous Wound Healing. Histol. Histopathol. 2008, 23, 1399–1407. [Google Scholar] [CrossRef]
  132. Broughton, G.; Janis, J.E.; Attinger, C.E. The Basic Science of Wound Healing. Plast. Reconstr. Surg. 2006, 117, 12S–34S. [Google Scholar] [CrossRef]
  133. Abkowitz, J.L.; Robinson, A.E.; Kale, S.; Long, M.W.; Chen, J. Mobilization of Hematopoietic Stem Cells during Homeostasis and after Cytokine Exposure. Blood 2003, 102, 1249–1253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Zarbock, A.; Ley, K.; McEver, R.P.; Hidalgo, A. Leukocyte Ligands for Endothelial Selectins: Specialized Glycoconjugates That Mediate Rolling and Signaling under Flow. Blood 2011, 118, 6743–6751. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Rahman, A.; Kefer, J.; Bando, M.; Niles, W.D.; Malik, A.B. E-Selectin Expression in Human Endothelial Cells by TNF-Alpha-Induced Oxidant Generation and NF-KappaB Activation. Am. J. Physiol. 1998, 275, L533–L544. [Google Scholar] [CrossRef] [PubMed]
  136. Ley, K.; Laudanna, C.; Cybulsky, M.I.; Nourshargh, S. Getting to the Site of Inflammation: The Leukocyte Adhesion Cascade Updated. Nat. Rev. Immunol. 2007, 7, 678–689. [Google Scholar] [CrossRef]
  137. Mezu-Ndubuisi, O.J.; Maheshwari, A. The Role of Integrins in Inflammation and Angiogenesis. Pediatr. Res. 2021, 89, 1619–1626. [Google Scholar] [CrossRef]
  138. 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]
  139. Herter, J.; Zarbock, A. Integrin Regulation during Leukocyte Recruitment. J. Immunol. 2013, 190, 4451–4457. [Google Scholar] [CrossRef] [Green Version]
  140. McEver, R.P.; Zhu, C. Rolling Cell Adhesion. Annu. Rev. Cell Dev. Biol. 2010, 26, 363–396. [Google Scholar] [CrossRef]
  141. Kuijper, P.H.; Gallardo Torres, H.I.; Houben, L.A.; Lammers, J.W.; Zwaginga, J.J.; Koenderman, L. P-Selectin and MAC-1 Mediate Monocyte Rolling and Adhesion to ECM-Bound Platelets under Flow Conditions. J. Leukoc. Biol. 1998, 64, 467–473. [Google Scholar] [CrossRef]
  142. Wolf, D.; Anto-Michel, N.; Blankenbach, H.; Wiedemann, A.; Buscher, K.; Hohmann, J.D.; Lim, B.; Bäuml, M.; Marki, A.; Mauler, M.; et al. A Ligand-Specific Blockade of the Integrin Mac-1 Selectively Targets Pathologic Inflammation While Maintaining Protective Host-Defense. Nat. Commun. 2018, 9, 525. [Google Scholar] [CrossRef] [Green Version]
  143. Detmers, P.A.; Lo, S.K.; Olsen-Egbert, E.; Walz, A.; Baggiolini, M.; Cohn, Z.A. Neutrophil-Activating Protein 1/Interleukin 8 Stimulates the Binding Activity of the Leukocyte Adhesion Receptor CD11b/CD18 on Human Neutrophils. J. Exp. Med. 1990, 171, 1155–1162. [Google Scholar] [CrossRef] [PubMed]
  144. 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]
  145. Wang, S.; Voisin, M.-B.; Larbi, K.Y.; Dangerfield, J.; Scheiermann, C.; Tran, M.; Maxwell, P.H.; Sorokin, L.; Nourshargh, S. Venular Basement Membranes Contain Specific Matrix Protein Low Expression Regions That Act as Exit Points for Emigrating Neutrophils. J. Exp. Med. 2006, 203, 1519–1532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Engelhardt, E.; Toksoy, A.; Goebeler, M.; Debus, S.; Bröcker, E.B.; Gillitzer, R. Chemokines IL-8, GROalpha, MCP-1, IP-10, and Mig Are Sequentially and Differentially Expressed during Phase-Specific Infiltration of Leukocyte Subsets in Human Wound Healing. Am. J. Pathol. 1998, 153, 1849–1860. [Google Scholar] [CrossRef]
  147. Ng, M.F.Y. The Role of Mast Cells in Wound Healing. Int. Wound J. 2010, 7, 55–61. [Google Scholar] [CrossRef]
  148. Younan, G.; Suber, F.; Xing, W.; Shi, T.; Kunori, Y.; Abrink, M.; Pejler, G.; Schlenner, S.M.; Rodewald, H.-R.; Moore, F.D.; et al. The Inflammatory Response after an Epidermal Burn Depends on the Activities of Mouse Mast Cell Proteases 4 and 5. J. Immunol. 2010, 185, 7681–7690. [Google Scholar] [CrossRef] [Green Version]
  149. Kolaczkowska, E.; Kubes, P. Neutrophil Recruitment and Function in Health and Inflammation. Nat. Rev. Immunol. 2013, 13, 159–175. [Google Scholar] [CrossRef]
  150. De Oliveira, S.; Rosowski, E.E.; Huttenlocher, A. Neutrophil Migration in Infection and Wound Repair: Going Forward in Reverse. Nat. Rev. Immunol. 2016, 16, 378–391. [Google Scholar] [CrossRef] [Green Version]
  151. Gillitzer, R.; Goebeler, M. Chemokines in Cutaneous Wound Healing. J. Leukoc. Biol. 2001, 69, 513–521. [Google Scholar] [CrossRef]
  152. Su, Y.; Richmond, A. Chemokine Regulation of Neutrophil Infiltration of Skin Wounds. Adv. Wound Care 2015, 4, 631–640. [Google Scholar] [CrossRef] [Green Version]
  153. Lämmermann, T. In the Eye of the Neutrophil Swarm-Navigation Signals That Bring Neutrophils Together in Inflamed and Infected Tissues. J. Leukoc. Biol. 2016, 100, 55–63. [Google Scholar] [CrossRef] [PubMed]
  154. Ng, L.G.; Qin, J.S.; Roediger, B.; Wang, Y.; Jain, R.; Cavanagh, L.L.; Smith, A.L.; Jones, C.A.; de Veer, M.; Grimbaldeston, M.A.; et al. Visualizing the Neutrophil Response to Sterile Tissue Injury in Mouse Dermis Reveals a Three-Phase Cascade of Events. J. Investig. Dermatol. 2011, 131, 2058–2068. [Google Scholar] [CrossRef] [Green Version]
  155. Brinkmann, V.; Reichard, U.; Goosmann, C.; Fauler, B.; Uhlemann, Y.; Weiss, D.S.; Weinrauch, Y.; Zychlinsky, A. Neutrophil Extracellular Traps Kill Bacteria. Science 2004, 303, 1532–1535. [Google Scholar] [CrossRef] [PubMed]
  156. Ellis, S.; Lin, E.J.; Tartar, D. Immunology of Wound Healing. Curr. Dermatol. Rep. 2018, 7, 350–358. [Google Scholar] [CrossRef] [PubMed]
  157. Moor, A.N.; Vachon, D.J.; Gould, L.J. Proteolytic Activity in Wound Fluids and Tissues Derived from Chronic Venous Leg Ulcers. Wound Repair Regen. 2009, 17, 832–839. [Google Scholar] [CrossRef] [PubMed]
  158. Degradation of the Epidermal-Dermal Junction by Proteolytic Enzymes from Human Skin and Human Polymorphonuclear Leukocytes. J. Exp. Med. 1984, 160, 1027–1042. [CrossRef] [Green Version]
  159. Pirilä, E.; Korpi, J.T.; Korkiamäki, T.; Jahkola, T.; Gutierrez-Fernandez, A.; Lopez-Otin, C.; Saarialho-Kere, U.; Salo, T.; Sorsa, T. Collagenase-2 (MMP-8) and Matrilysin-2 (MMP-26) Expression in Human Wounds of Different Etiologies. Wound Repair Regen. 2007, 15, 47–57. [Google Scholar] [CrossRef]
  160. Butin-Israeli, V.; Bui, T.M.; Wiesolek, H.L.; Mascarenhas, L.; Lee, J.J.; Mehl, L.C.; Knutson, K.R.; Adam, S.A.; Goldman, R.D.; Beyder, A.; et al. Neutrophil-Induced Genomic Instability Impedes Resolution of Inflammation and Wound Healing. J. Clin. Investig. 2019, 129, 712–726. [Google Scholar] [CrossRef] [Green Version]
  161. Levin, R.; Grinstein, S.; Canton, J. The Life Cycle of Phagosomes: Formation, Maturation, and Resolution. Immunol. Rev. 2016, 273, 156–179. [Google Scholar] [CrossRef]
  162. Lee, W.L.; Harrison, R.E.; Grinstein, S. Phagocytosis by Neutrophils. Microbes Infect. 2003, 5, 1299–1306. [Google Scholar] [CrossRef]
  163. Borregaard, N.; Cowland, J.B. Granules of the Human Neutrophilic Polymorphonuclear Leukocyte. Blood 1997, 89, 3503–3521. [Google Scholar] [CrossRef] [PubMed]
  164. Reeves, E.P.; Lu, H.; Jacobs, H.L.; Messina, C.G.M.; Bolsover, S.; Gabella, G.; Potma, E.O.; Warley, A.; Roes, J.; Segal, A.W. Killing Activity of Neutrophils Is Mediated through Activation of Proteases by K+ Flux. Nature 2002, 416, 291–297. [Google Scholar] [CrossRef] [PubMed]
  165. Segel, G.B.; Halterman, M.W.; Lichtman, M.A. The Paradox of the Neutrophil’s Role in Tissue Injury. J. Leukoc. Biol. 2011, 89, 359–372. [Google Scholar] [CrossRef] [PubMed]
  166. Gramegna, A.; Amati, F.; Terranova, L.; Sotgiu, G.; Tarsia, P.; Miglietta, D.; Calderazzo, M.A.; Aliberti, S.; Blasi, F. Neutrophil Elastase in Bronchiectasis. Respir. Res. 2017, 18, 211. [Google Scholar] [CrossRef] [PubMed]
  167. Owen, C.A.; Campbell, M.A.; Sannes, P.L.; Boukedes, S.S.; Campbell, E.J. Cell Surface-Bound Elastase and Cathepsin G on Human Neutrophils: A Novel, Non-Oxidative Mechanism by Which Neutrophils Focus and Preserve Catalytic Activity of Serine Proteinases. J. Cell Biol. 1995, 131, 775–789. [Google Scholar] [CrossRef]
  168. Sagel, S.D.; Wagner, B.D.; Anthony, M.M.; Emmett, P.; Zemanick, E.T. Sputum Biomarkers of Inflammation and Lung Function Decline in Children with Cystic Fibrosis. Am. J. Respir. Crit. Care Med. 2012, 186, 857–865. [Google Scholar] [CrossRef] [Green Version]
  169. Döring, G.; Frank, F.; Boudier, C.; Herbert, S.; Fleischer, B.; Bellon, G. Cleavage of Lymphocyte Surface Antigens CD2, CD4, and CD8 by Polymorphonuclear Leukocyte Elastase and Cathepsin G in Patients with Cystic Fibrosis. J. Immunol. 1995, 154, 4842–4850. [Google Scholar]
  170. Ma, Y.; Yabluchanskiy, A.; Iyer, R.P.; Cannon, P.L.; Flynn, E.R.; Jung, M.; Henry, J.; Cates, C.A.; Deleon-Pennell, K.Y.; Lindsey, M.L. Temporal Neutrophil Polarization Following Myocardial Infarction. Cardiovasc. Res. 2016, 110, 51–61. [Google Scholar] [CrossRef] [Green Version]
  171. Elliott, M.R.; Koster, K.M.; Murphy, P.S. Efferocytosis Signaling in the Regulation of Macrophage Inflammatory Responses. J. Immunol. 2017, 198, 1387–1394. [Google Scholar] [CrossRef] [Green Version]
  172. Havixbeck, J.J.; Rieger, A.M.; Wong, M.E.; Hodgkinson, J.W.; Barreda, D.R. Neutrophil Contributions to the Induction and Regulation of the Acute Inflammatory Response in Teleost Fish. J. Leukoc. Biol. 2016, 99, 241–252. [Google Scholar] [CrossRef] [Green Version]
  173. Devalaraja, R.M.; Nanney, L.B.; Du, J.; Qian, Q.; Yu, Y.; Devalaraja, M.N.; Richmond, A. Delayed Wound Healing in CXCR2 Knockout Mice. J. Investig. Dermatol. 2000, 115, 234–244. [Google Scholar] [CrossRef] [Green Version]
  174. Nishio, N.; Okawa, Y.; Sakurai, H.; Isobe, K. Neutrophil Depletion Delays Wound Repair in Aged Mice. Age 2008, 30, 11–19. [Google Scholar] [CrossRef] [Green Version]
  175. Ortmann, W.; Kolaczkowska, E. Age Is the Work of Art? Impact of Neutrophil and Organism Age on Neutrophil Extracellular Trap Formation. Cell Tissue Res. 2018, 371, 473–488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Tseng, C.W.; Liu, G.Y. Expanding Roles of Neutrophils in Aging Hosts. Curr. Opin. Immunol. 2014, 29, 43–48. [Google Scholar] [CrossRef] [PubMed]
  177. Dovi, J.V.; He, L.-K.; DiPietro, L.A. Accelerated Wound Closure in Neutrophil-Depleted Mice. J. Leukoc. Biol. 2003, 73, 448–455. [Google Scholar] [CrossRef] [PubMed]
  178. Minutti, C.M.; Knipper, J.A.; Allen, J.E.; Zaiss, D.M.W. Tissue-Specific Contribution of Macrophages to Wound Healing. Semin. Cell Dev. Biol. 2017, 61, 3–11. [Google Scholar] [CrossRef] [Green Version]
  179. Yanez, D.A.; Lacher, R.K.; Vidyarthi, A.; Colegio, O.R. The Role of Macrophages in Skin Homeostasis. Pflügers Arch.-Eur. J. Physiol. 2017, 469, 455–463. [Google Scholar] [CrossRef] [Green Version]
  180. Xuan, W.; Qu, Q.; Zheng, B.; Xiong, S.; Fan, G.-H. The Chemotaxis of M1 and M2 Macrophages Is Regulated by Different Chemokines. J. Leukoc. Biol. 2015, 97, 61–69. [Google Scholar] [CrossRef]
  181. DiPietro, L.A.; Polverini, P.J.; Rahbe, S.M.; Kovacs, E.J. Modulation of JE/MCP-1 Expression in Dermal Wound Repair. Am. J. Pathol. 1995, 146, 868–875. [Google Scholar]
  182. Gordon, S. Alternative Activation of Macrophages. Nat. Rev. Immunol. 2003, 3, 23–35. [Google Scholar] [CrossRef]
  183. Song, E.; Ouyang, N.; Hörbelt, M.; Antus, B.; Wang, M.; Exton, M.S. Influence of Alternatively and Classically Activated Macrophages on Fibrogenic Activities of Human Fibroblasts. Cell. Immunol. 2000, 204, 19–28. [Google Scholar] [CrossRef] [PubMed]
  184. Novak, M.L.; Koh, T.J. Macrophage Phenotypes during Tissue Repair. J. Leukoc. Biol. 2013, 93, 875–881. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Murray, P.J.; Allen, J.E.; Biswas, S.K.; Fisher, E.A.; Gilroy, D.W.; Goerdt, S.; Gordon, S.; Hamilton, J.A.; Ivashkiv, L.B.; Lawrence, T.; et al. Macrophage Activation and Polarization: Nomenclature and Experimental Guidelines. Immunity 2014, 41, 14–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Stout, R.D.; Jiang, C.; Matta, B.; Tietzel, I.; Watkins, S.K.; Suttles, J. Macrophages Sequentially Change Their Functional Phenotype in Response to Changes in Microenvironmental Influences. J. Immunol. 2005, 175, 342–349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Slauch, J.M. How Does the Oxidative Burst of Macrophages Kill Bacteria? Still an Open Question. Mol. Microbiol. 2011, 80, 580–583. [Google Scholar] [CrossRef] [Green Version]
  188. Kim, S.Y.; Nair, M.G. Macrophages in Wound Healing: Activation and Plasticity. Immunol. Cell Biol. 2019, 97, 258–267. [Google Scholar] [CrossRef]
  189. Hart, P.H.; Jones, C.A.; Finlay-Jones, J.J. Monocytes Cultured in Cytokine-Defined Environments Differ from Freshly Isolated Monocytes in Their Responses to IL-4 and IL-10. J. Leukoc. Biol. 1995, 57, 909–918. [Google Scholar] [CrossRef]
  190. Goren, I.; Allmann, N.; Yogev, N.; Schürmann, C.; Linke, A.; Holdener, M.; Waisman, A.; Pfeilschifter, J.; Frank, S. A Transgenic Mouse Model of Inducible Macrophage Depletion: Effects of Diphtheria Toxin-Driven Lysozyme M-Specific Cell Lineage Ablation on Wound Inflammatory, Angiogenic, and Contractive Processes. Am. J. Pathol. 2009, 175, 132–147. [Google Scholar] [CrossRef] [Green Version]
  191. Mirza, R.; DiPietro, L.A.; Koh, T.J. Selective and Specific Macrophage Ablation Is Detrimental to Wound Healing in Mice. Am. J. Pathol. 2009, 175, 2454–2462. [Google Scholar] [CrossRef] [Green Version]
  192. Zhu, Z.; Ding, J.; Ma, Z.; Iwashina, T.; Tredget, E.E. Systemic Depletion of Macrophages in the Subacute Phase of Wound Healing Reduces Hypertrophic Scar Formation. Wound Repair Regen. 2016, 24, 644–656. [Google Scholar] [CrossRef]
  193. Koh, T.J.; DiPietro, L.A. Inflammation and Wound Healing: The Role of the Macrophage. Expert Rev. Mol. Med. 2011, 13, e23. [Google Scholar] [CrossRef] [Green Version]
  194. Kreisel, D.; Nava, R.G.; Li, W.; Zinselmeyer, B.H.; Wang, B.; Lai, J.; Pless, R.; Gelman, A.E.; Krupnick, A.S.; Miller, M.J. In Vivo Two-Photon Imaging Reveals Monocyte-Dependent Neutrophil Extravasation during Pulmonary Inflammation. Proc. Natl. Acad. Sci. USA 2010, 107, 18073–18078. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Sorokin, L. The Impact of the Extracellular Matrix on Inflammation. Nat. Rev. Immunol. 2010, 10, 712–723. [Google Scholar] [CrossRef] [PubMed]
  196. Davies, L.C.; Jenkins, S.J.; Allen, J.E.; Taylor, P.R. Tissue-Resident Macrophages. Nat. Immunol. 2013, 14, 986–995. [Google Scholar] [CrossRef] [PubMed]
  197. Ferrante, C.J.; Leibovich, S.J. Regulation of Macrophage Polarization and Wound Healing. Adv. Wound Care 2012, 1, 10–16. [Google Scholar] [CrossRef] [Green Version]
  198. Ogle, M.E.; Segar, C.E.; Sridhar, S.; Botchwey, E.A. Monocytes and Macrophages in Tissue Repair: Implications for Immunoregenerative Biomaterial Design. Exp. Biol. Med. 2016, 241, 1084–1097. [Google Scholar] [CrossRef]
  199. Leitinger, N.; Schulman, I.G. Phenotypic Polarization of Macrophages in Atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 2013, 33, 1120–1126. [Google Scholar] [CrossRef] [Green Version]
  200. Spiller, K.L.; Anfang, R.R.; Spiller, K.J.; Ng, J.; Nakazawa, K.R.; Daulton, J.W.; Vunjak-Novakovic, G. The Role of Macrophage Phenotype in Vascularization of Tissue Engineering Scaffolds. Biomaterials 2014, 35, 4477–4488. [Google Scholar] [CrossRef] [Green Version]
  201. Filardy, A.A.; Pires, D.R.; Nunes, M.P.; Takiya, C.M.; Freire-de-Lima, C.G.; Ribeiro-Gomes, F.L.; DosReis, G.A. Proinflammatory Clearance of Apoptotic Neutrophils Induces an IL-12lowIL-10high Regulatory Phenotype in Macrophages. J. Immunol. 2010, 185, 2044–2050. [Google Scholar] [CrossRef] [Green Version]
  202. Hesketh, M.; Sahin, K.B.; West, Z.E.; Murray, R.Z. Macrophage Phenotypes Regulate Scar Formation and Chronic Wound Healing. Int. J. Mol. Sci. 2017, 18, 1545. [Google Scholar] [CrossRef] [Green Version]
  203. Rőszer, T. Understanding the Mysterious M2 Macrophage through Activation Markers and Effector Mechanisms. Mediat. Inflamm. 2015, 2015, 816460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Zizzo, G.; Hilliard, B.A.; Monestier, M.; Cohen, P.L. Efficient Clearance of Early Apoptotic Cells by Human Macrophages Requires M2c Polarization and MerTK Induction. J. Immunol. 2012, 189, 3508–3520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Wang, Q.; Ni, H.; Lan, L.; Wei, X.; Xiang, R.; Wang, Y. Fra-1 Protooncogene Regulates IL-6 Expression in Macrophages and Promotes the Generation of M2d Macrophages. Cell Res. 2010, 20, 701–712. [Google Scholar] [CrossRef] [PubMed]
  206. Duluc, D.; Delneste, Y.; Tan, F.; Moles, M.-P.; Grimaud, L.; Lenoir, J.; Preisser, L.; Anegon, I.; Catala, L.; Ifrah, N.; et al. Tumor-Associated Leukemia Inhibitory Factor and IL-6 Skew Monocyte Differentiation into Tumor-Associated Macrophage-like Cells. Blood 2007, 110, 4319–4330. [Google Scholar] [CrossRef]
  207. Atri, C.; Guerfali, F.Z.; Laouini, D. Role of Human Macrophage Polarization in Inflammation during Infectious Diseases. Int. J. Mol. Sci. 2018, 19, 1801. [Google Scholar] [CrossRef] [Green Version]
  208. Shapouri-Moghaddam, A.; Mohammadian, S.; Vazini, H.; Taghadosi, M.; Esmaeili, S.-A.; Mardani, F.; Seifi, B.; Mohammadi, A.; Afshari, J.T.; Sahebkar, A. Macrophage Plasticity, Polarization, and Function in Health and Disease. J. Cell Physiol. 2018, 233, 6425–6440. [Google Scholar] [CrossRef] [PubMed]
  209. Steinman, R.M.; Cohn, Z.A. Identification of a Novel Cell Type in Peripheral Lymphoid Organs of Mice. I. Morphology, Quantitation, Tissue Distribution. J. Exp. Med. 1973, 137, 1142–1162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Steinman, R.M.; Gutchinov, B.; Witmer, M.D.; Nussenzweig, M.C. Dendritic Cells Are the Principal Stimulators of the Primary Mixed Leukocyte Reaction in Mice. J. Exp. Med. 1983, 157, 613–627. [Google Scholar] [CrossRef] [Green Version]
  211. Steinman, R.M.; Banchereau, J. Taking Dendritic Cells into Medicine. Nature 2007, 449, 419–426. [Google Scholar] [CrossRef] [PubMed]
  212. Balan, S.; Saxena, M.; Bhardwaj, N. Dendritic Cell Subsets and Locations. Int. Rev. Cell Mol. Biol. 2019, 348, 1–68. [Google Scholar] [CrossRef]
  213. Gregorio, J.; Meller, S.; Conrad, C.; Di Nardo, A.; Homey, B.; Lauerma, A.; Arai, N.; Gallo, R.L.; Digiovanni, J.; Gilliet, M. Plasmacytoid Dendritic Cells Sense Skin Injury and Promote Wound Healing through Type I Interferons. J. Exp. Med. 2010, 207, 2921–2930. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Jegalian, A.G.; Facchetti, F.; Jaffe, E.S. Plasmacytoid Dendritic Cells: Physiologic Roles and Pathologic States. Adv. Anat. Pathol. 2009, 16, 392–404. [Google Scholar] [CrossRef] [PubMed]
  215. McLachlan, J.B.; Catron, D.M.; Moon, J.J.; Jenkins, M.K. Dendritic Cell Antigen Presentation Drives Simultaneous Cytokine Production by Effector and Regulatory T Cells in Inflamed Skin. Immunity 2009, 30, 277–288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Vinish, M.; Cui, W.; Stafford, E.; Bae, L.; Hawkins, H.; Cox, R.; Toliver-Kinsky, T. Dendritic Cells Modulate Burn Wound Healing by Enhancing Early Proliferation. Wound Repair Regen. 2016, 24, 6–13. [Google Scholar] [CrossRef] [Green Version]
  217. Gao, N.; Yin, J.; Yoon, G.S.; Mi, Q.-S.; Yu, F.-S.X. Dendritic Cell–Epithelium Interplay Is a Determinant Factor for Corneal Epithelial Wound Repair. Am. J. Pathol. 2011, 179, 2243–2253. [Google Scholar] [CrossRef]
  218. Romani, N.; Holzmann, S.; Tripp, C.H.; Koch, F.; Stoitzner, P. Langerhans Cells—Dendritic Cells of the Epidermis. APMIS 2003, 111, 725–740. [Google Scholar] [CrossRef]
  219. Kubo, A.; Nagao, K.; Yokouchi, M.; Sasaki, H.; Amagai, M. External Antigen Uptake by Langerhans Cells with Reorganization of Epidermal Tight Junction Barriers. J. Exp. Med. 2009, 206, 2937–2946. [Google Scholar] [CrossRef] [Green Version]
  220. Joffre, O.; Nolte, M.A.; Spörri, R.; Reis e Sousa, C. Inflammatory Signals in Dendritic Cell Activation and the Induction of Adaptive Immunity. Immunol. Rev. 2009, 227, 234–247. [Google Scholar] [CrossRef]
  221. Tang, A.; Amagai, M.; Granger, L.G.; Stanley, J.R.; Udey, M.C. Adhesion of Epidermal Langerhans Cells to Keratinocytes Mediated by E-Cadherin. Nature 1993, 361, 82–85. [Google Scholar] [CrossRef]
  222. Ratzinger, G.; Stoitzner, P.; Ebner, S.; Lutz, M.B.; Layton, G.T.; Rainer, C.; Senior, R.M.; Shipley, J.M.; Fritsch, P.; Schuler, G.; et al. Matrix Metalloproteinases 9 and 2 Are Necessary for the Migration of Langerhans Cells and Dermal Dendritic Cells from Human and Murine Skin. J. Immunol. 2002, 168, 4361–4371. [Google Scholar] [CrossRef] [Green Version]
  223. Kabashima, K.; Shiraishi, N.; Sugita, K.; Mori, T.; Onoue, A.; Kobayashi, M.; Sakabe, J.-I.; Yoshiki, R.; Tamamura, H.; Fujii, N.; et al. CXCL12-CXCR4 Engagement Is Required for Migration of Cutaneous Dendritic Cells. Am. J. Pathol. 2007, 171, 1249–1257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Seneschal, J.; Clark, R.A.; Gehad, A.; Baecher-Allan, C.M.; Kupper, T.S. Human Epidermal Langerhans Cells Maintain Immune Homeostasis in Skin by Activating Skin Resident Regulatory T Cells. Immunity 2012, 36, 873–884. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. West, H.C.; Bennett, C.L. Redefining the Role of Langerhans Cells as Immune Regulators within the Skin. Front. Immunol. 2018, 8, 1941. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Stojadinovic, O.; Yin, N.; Lehmann, J.; Pastar, I.; Kirsner, R.S.; Tomic-Canic, M. Increased Number of Langerhans Cells in the Epidermis of Diabetic Foot Ulcers Correlates with Healing Outcome. Immunol. Res. 2013, 57, 222–228. [Google Scholar] [CrossRef] [Green Version]
  227. Sonoda, T.; Kitamura, Y.; Haku, Y.; Hara, H.; Mori, K.J. Mast-Cell Precursors in Various Haematopoietic Colonies of Mice Produced in Vivo and in Vitro. Br. J. Haematol. 1983, 53, 611–620. [Google Scholar] [CrossRef]
  228. Artuc, M.; Hermes, B.; Steckelings, U.M.; Grützkau, A.; Henz, B.M. Mast Cells and Their Mediators in Cutaneous Wound Healing—Active Participants or Innocent Bystanders? Exp. Dermatol. 1999, 8, 1–16. [Google Scholar] [CrossRef]
  229. Trautmann, A.; Toksoy, A.; Engelhardt, E.; Bröcker, E.B.; Gillitzer, R. Mast Cell Involvement in Normal Human Skin Wound Healing: Expression of Monocyte Chemoattractant Protein-1 Is Correlated with Recruitment of Mast Cells Which Synthesize Interleukin-4 in Vivo. J. Pathol. 2000, 190, 100–106. [Google Scholar] [CrossRef]
  230. Mukai, K.; Tsai, M.; Saito, H.; Galli, S.J. Mast Cells as Sources of Cytokines, Chemokines, and Growth Factors. Immunol. Rev. 2018, 282, 121–150. [Google Scholar] [CrossRef]
  231. Elieh Ali Komi, D.; Bjermer, L. Mast Cell-Mediated Orchestration of the Immune Responses in Human Allergic Asthma: Current Insights. Clin. Rev. Allergy Immunol. 2019, 56, 234–247. [Google Scholar] [CrossRef] [Green Version]
  232. Komi, D.E.A.; Rambasek, T.; Wöhrl, S. Mastocytosis: From a Molecular Point of View. Clin. Rev. Allergy Immunol. 2018, 54, 397–411. [Google Scholar] [CrossRef]
  233. Elieh Ali Komi, D.; Rambasek, T.; Bielory, L. Clinical Implications of Mast Cell Involvement in Allergic Conjunctivitis. Allergy 2018, 73, 528–539. [Google Scholar] [CrossRef] [PubMed]
  234. Wilgus, T.A.; Wulff, B.C. The Importance of Mast Cells in Dermal Scarring. Adv. Wound Care 2014, 3, 356–365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Komi, D.E.A.; Khomtchouk, K.; Santa Maria, P.L. A Review of the Contribution of Mast Cells in Wound Healing: Involved Molecular and Cellular Mechanisms. Clin. Rev. Allergy Immunol. 2020, 58, 298–312. [Google Scholar] [CrossRef] [PubMed]
  236. Rao, K.N.; Brown, M.A. Mast Cells: Multifaceted Immune Cells with Diverse Roles in Health and Disease. Ann. N. Y. Acad. Sci. 2008, 1143, 83–104. [Google Scholar] [CrossRef]
  237. Nakano, T.; Sonoda, T.; Hayashi, C.; Yamatodani, A.; Kanayama, Y.; Yamamura, T.; Asai, H.; Yonezawa, T.; Kitamura, Y.; Galli, S.J. Fate of Bone Marrow-Derived Cultured Mast Cells after Intracutaneous, Intraperitoneal, and Intravenous Transfer into Genetically Mast Cell-Deficient W/Wv Mice. Evidence That Cultured Mast Cells Can Give Rise to Both Connective Tissue Type and Mucosal Mast Cells. J. Exp. Med. 1985, 162, 1025–1043. [Google Scholar] [CrossRef] [Green Version]
  238. Wojta, J.; Kaun, C.; Zorn, G.; Ghannadan, M.; Hauswirth, A.W.; Sperr, W.R.; Fritsch, G.; Printz, D.; Binder, B.R.; Schatzl, G.; et al. C5a Stimulates Production of Plasminogen Activator Inhibitor-1 in Human Mast Cells and Basophils. Blood 2002, 100, 517–523. [Google Scholar] [CrossRef]
  239. Oschatz, C.; Maas, C.; Lecher, B.; Jansen, T.; Björkqvist, J.; Tradler, T.; Sedlmeier, R.; Burfeind, P.; Cichon, S.; Hammerschmidt, S.; et al. Mast Cells Increase Vascular Permeability by Heparin-Initiated Bradykinin Formation in Vivo. Immunity 2011, 34, 258–268. [Google Scholar] [CrossRef]
  240. Kennelly, R.; Conneely, J.B.; Bouchier-Hayes, D.; Winter, D.C. Mast Cells in Tissue Healing: From Skin to the Gastrointestinal Tract. Curr. Pharm. Des. 2011, 17, 3772–3775. [Google Scholar] [CrossRef]
  241. Theoharides, T.C.; Alysandratos, K.-D.; Angelidou, A.; Delivanis, D.-A.; Sismanopoulos, N.; Zhang, B.; Asadi, S.; Vasiadi, M.; Weng, Z.; Miniati, A.; et al. Mast Cells and Inflammation. Biochim. Biophys. Acta 2012, 1822, 21–33. [Google Scholar] [CrossRef] [Green Version]
  242. Wan, S.-W.; Wu-Hsieh, B.A.; Lin, Y.-S.; Chen, W.-Y.; Huang, Y.; Anderson, R. The Monocyte-Macrophage-Mast Cell Axis in Dengue Pathogenesis. J. Biomed. Sci. 2018, 25, 77. [Google Scholar] [CrossRef] [Green Version]
  243. Egozi, E.I.; Ferreira, A.M.; Burns, A.L.; Gamelli, R.L.; Dipietro, L.A. Mast Cells Modulate the Inflammatory but Not the Proliferative Response in Healing Wounds. Wound Repair Regen. 2003, 11, 46–54. [Google Scholar] [CrossRef]
  244. Iba, Y.; Shibata, A.; Kato, M.; Masukawa, T. Possible Involvement of Mast Cells in Collagen Remodeling in the Late Phase of Cutaneous Wound Healing in Mice. Int. Immunopharmacol. 2004, 4, 1873–1880. [Google Scholar] [CrossRef]
  245. Ishikawa, T.; Kanda, N.; Hau, C.S.; Tada, Y.; Watanabe, S. Histamine Induces Human Beta-Defensin-3 Production in Human Keratinocytes. J. Dermatol. Sci. 2009, 56, 121–127. [Google Scholar] [CrossRef] [PubMed]
  246. Huttunen, M.; Aalto, M.L.; Harvima, R.J.; Horsmanheimo, M.; Harvima, I.T. Alterations in Mast Cells Showing Tryptase and Chymase Activity in Epithelializating and Chronic Wounds. Exp. Dermatol. 2000, 9, 258–265. [Google Scholar] [CrossRef]
  247. Sivamani, R.K. Eicosanoids and Keratinocytes in Wound Healing. Adv. Wound Care 2014, 3, 476–481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Huttunen, M.; Hyttinen, M.; Nilsson, G.; Butterfield, J.H.; Horsmanheimo, M.; Harvima, I.T. Inhibition of Keratinocyte Growth in Cell Culture and Whole Skin Culture by Mast Cell Mediators. Exp. Dermatol. 2001, 10, 184–192. [Google Scholar] [CrossRef] [PubMed]
  249. Tellechea, A.; Leal, E.C.; Kafanas, A.; Auster, M.E.; Kuchibhotla, S.; Ostrovsky, Y.; Tecilazich, F.; Baltzis, D.; Zheng, Y.; Carvalho, E.; et al. Mast Cells Regulate Wound Healing in Diabetes. Diabetes 2016, 65, 2006–2019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  250. Maltby, S.; Khazaie, K.; McNagny, K.M. Mast Cells in Tumor Growth: Angiogenesis, Tissue Remodelling and Immune-Modulation. Biochim. Biophys. Acta 2009, 1796, 19–26. [Google Scholar] [CrossRef] [Green Version]
  251. Jameson, J.; Ugarte, K.; Chen, N.; Yachi, P.; Fuchs, E.; Boismenu, R.; Havran, W.L. A Role for Skin Gammadelta T Cells in Wound Repair. Science 2002, 296, 747–749. [Google Scholar] [CrossRef]
  252. Jameson, J.M.; Cauvi, G.; Witherden, D.A.; Havran, W.L. A Keratinocyte-Responsive Gamma Delta TCR Is Necessary for Dendritic Epidermal T Cell Activation by Damaged Keratinocytes and Maintenance in the Epidermis. J. Immunol. 2004, 172, 3573–3579. [Google Scholar] [CrossRef] [Green Version]
  253. Jameson, J.M.; Cauvi, G.; Sharp, L.L.; Witherden, D.A.; Havran, W.L. Gammadelta T Cell-Induced Hyaluronan Production by Epithelial Cells Regulates Inflammation. J. Exp. Med. 2005, 201, 1269–1279. [Google Scholar] [CrossRef] [Green Version]
  254. Keyes, B.E.; Liu, S.; Asare, A.; Naik, S.; Levorse, J.; Polak, L.; Lu, C.P.; Nikolova, M.; Pasolli, H.A.; Fuchs, E. Impaired Epidermal to Dendritic T Cell Signaling Slows Wound Repair in Aged Skin. Cell 2016, 167, 1323–1338.e14. [Google Scholar] [CrossRef] [Green Version]
  255. Witherden, D.A.; Watanabe, M.; Garijo, O.; Rieder, S.E.; Sarkisyan, G.; Cronin, S.J.F.; Verdino, P.; Wilson, I.A.; Kumanogoh, A.; Kikutani, H.; et al. The CD100 Receptor Interacts with Its Plexin B2 Ligand to Regulate Epidermal Γδ T Cell Function. Immunity 2012, 37, 314–325. [Google Scholar] [CrossRef] [Green Version]
  256. Havran, W.L.; Jameson, J.M. Epidermal T Cells and Wound Healing. J. Immunol. 2010, 184, 5423–5428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Panduro, M.; Benoist, C.; Mathis, D. Tissue Tregs. Annu. Rev. Immunol. 2016, 34, 609–633. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  258. Rubtsov, Y.P.; Rasmussen, J.P.; Chi, E.Y.; Fontenot, J.; Castelli, L.; Ye, X.; Treuting, P.; Siewe, L.; Roers, A.; Henderson, W.R.; et al. Regulatory T Cell-Derived Interleukin-10 Limits Inflammation at Environmental Interfaces. Immunity 2008, 28, 546–558. [Google Scholar] [CrossRef] [PubMed]
  259. Josefowicz, S.Z.; Lu, L.-F.; Rudensky, A.Y. Regulatory T Cells: Mechanisms of Differentiation and Function. Annu. Rev. Immunol. 2012, 30, 531–564. [Google Scholar] [CrossRef] [PubMed]
  260. Scharschmidt, T.C.; Vasquez, K.S.; Truong, H.-A.; Gearty, S.V.; Pauli, M.L.; Nosbaum, A.; Gratz, I.K.; Otto, M.; Moon, J.J.; Liese, J.; et al. A Wave of Regulatory T Cells into Neonatal Skin Mediates Tolerance to Commensal Microbes. Immunity 2015, 43, 1011–1021. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  261. Ali, N.; Zirak, B.; Rodriguez, R.S.; Pauli, M.L.; Truong, H.-A.; Lai, K.; Ahn, R.; Corbin, K.; Lowe, M.M.; Scharschmidt, T.C.; et al. Regulatory T Cells in Skin Facilitate Epithelial Stem Cell Differentiation. Cell 2017, 169, 1119–1129.e11. [Google Scholar] [CrossRef] [Green Version]
  262. Sanchez Rodriguez, R.; Pauli, M.L.; Neuhaus, I.M.; Yu, S.S.; Arron, S.T.; Harris, H.W.; Yang, S.H.-Y.; Anthony, B.A.; Sverdrup, F.M.; Krow-Lucal, E.; et al. Memory Regulatory T Cells Reside in Human Skin. J. Clin. Investig. 2014, 124, 1027–1036. [Google Scholar] [CrossRef] [Green Version]
  263. Zaiss, D.M.; Minutti, C.M.; Knipper, J.A. Immune- and Non-Immune-Mediated Roles of Regulatory T-Cells during Wound Healing. Immunology 2019, 157, 190–197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Nosbaum, A.; Prevel, N.; Truong, H.A.; Mehta, P.; Ettinger, M.; Scharschmidt, T.C.; Ali, N.H.; Pauli, M.L.; Abbas, A.K.; Rosenblum, M.D. Cutting Edge: Regulatory T Cells Facilitate Cutaneous Wound Healing. J. Immunol. 2016, 196, 2010–2014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Boothby, I.C.; Cohen, J.N.; Rosenblum, M.D. Regulatory T Cells in Skin Injury: At the Crossroads of Tolerance and Tissue Repair. Sci. Immunol. 2020, 5, eaaz9631. [Google Scholar] [CrossRef] [PubMed]
  266. Tiemessen, M.M.; Jagger, A.L.; Evans, H.G.; van Herwijnen, M.J.C.; John, S.; Taams, L.S. CD4+CD25+Foxp3+ Regulatory T Cells Induce Alternative Activation of Human Monocytes/Macrophages. Proc. Natl. Acad. Sci. USA 2007, 104, 19446–19451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Proto, J.D.; Doran, A.C.; Gusarova, G.; Yurdagul, A.; Sozen, E.; Subramanian, M.; Islam, M.N.; Rymond, C.C.; Du, J.; Hook, J.; et al. Regulatory T Cells Promote Macrophage Efferocytosis during Inflammation Resolution. Immunity 2018, 49, 666–677.e6. [Google Scholar] [CrossRef] [Green Version]
  268. Headland, S.E.; Norling, L.V. The Resolution of Inflammation: Principles and Challenges. Semin. Immunol. 2015, 27, 149–160. [Google Scholar] [CrossRef]
  269. Fullerton, J.N.; Gilroy, D.W. Resolution of Inflammation: A New Therapeutic Frontier. Nat. Rev. Drug Discov. 2016, 15, 551–567. [Google Scholar] [CrossRef]
  270. Bratton, D.L.; Henson, P.M. Neutrophil Clearance: When the Party Is over, Clean-up Begins. Trends Immunol. 2011, 32, 350–357. [Google Scholar] [CrossRef] [Green Version]
  271. Jun, J.-I.; Kim, K.-H.; Lau, L.F. The Matricellular Protein CCN1 Mediates Neutrophil Efferocytosis in Cutaneous Wound Healing. Nat. Commun. 2015, 6, 7386. [Google Scholar] [CrossRef] [Green Version]
  272. Ji, J.; Fan, J. Neutrophil in Reverse Migration: Role in Sepsis. Front. Immunol. 2021, 12, 656039. [Google Scholar] [CrossRef]
  273. Chen, W.Y.J.; Rogers, A.A. Recent Insights into the Causes of Chronic Leg Ulceration in Venous Diseases and Implications on Other Types of Chronic Wounds. Wound Repair Regen. 2007, 15, 434–449. [Google Scholar] [CrossRef] [PubMed]
  274. Zhao, H.; Li, W.; Lu, Z.; Sheng, Z.; Yao, Y. The Growing Spectrum of Anti-Inflammatory Interleukins and Their Potential Roles in the Development of Sepsis. J. Interf. Cytokine Res. 2015, 35, 242–251. [Google Scholar] [CrossRef]
  275. Li, M.O.; Flavell, R.A. Contextual Regulation of Inflammation: A Duet by Transforming Growth Factor-Beta and Interleukin-10. Immunity 2008, 28, 468–476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  276. Chernoff, A.E.; Granowitz, E.V.; Shapiro, L.; Vannier, E.; Lonnemann, G.; Angel, J.B.; Kennedy, J.S.; Rabson, A.R.; Wolff, S.M.; Dinarello, C.A. A Randomized, Controlled Trial of IL-10 in Humans. Inhibition of Inflammatory Cytokine Production and Immune Responses. J. Immunol. 1995, 154, 5492–5499. [Google Scholar] [PubMed]
  277. Arango Duque, G.; Descoteaux, A. Macrophage Cytokines: Involvement in Immunity and Infectious Diseases. Front. Immunol. 2014, 5, 491. [Google Scholar] [CrossRef] [Green Version]
  278. Palolahti, M.; Lauharanta, J.; Stephens, R.W.; Kuusela, P.; Vaheri, A. Proteolytic Activity in Leg Ulcer Exudate. Exp. Dermatol. 1993, 2, 29–37. [Google Scholar] [CrossRef]
  279. Harris, I.R.; Yee, K.C.; Walters, C.E.; Cunliffe, W.J.; Kearney, J.N.; Wood, E.J.; Ingham, E. Cytokine and Protease Levels in Healing and Non-Healing Chronic Venous Leg Ulcers. Exp. Dermatol. 1995, 4, 342–349. [Google Scholar] [CrossRef]
  280. Barrick, B.; Campbell, E.J.; Owen, C.A. Leukocyte Proteinases in Wound Healing: Roles in Physiologic and Pathologic Processes. Wound Repair Regen. 1999, 7, 410–422. [Google Scholar] [CrossRef]
  281. Saarialho-Kere, U.K. Patterns of Matrix Metalloproteinase and TIMP Expression in Chronic Ulcers. Arch. Dermatol. Res. 1998, 290 (Suppl. S1), S47–S54. [Google Scholar] [CrossRef]
  282. Demidova-Rice, T.N.; Hamblin, M.R.; Herman, I.M. Acute and Impaired Wound Healing: Pathophysiology and Current Methods for Drug Delivery, Part 1: Normal and Chronic Wounds: Biology, Causes, and Approaches to Care. Adv. Skin Wound Care 2012, 25, 304–314. [Google Scholar] [CrossRef] [Green Version]
  283. Diegelmann, R.F.; Evans, M.C. Wound Healing: An Overview of Acute, Fibrotic and Delayed Healing. Front. Biosci. 2004, 9, 283–289. [Google Scholar] [CrossRef]
  284. Mast, B.A.; Schultz, G.S. Interactions of Cytokines, Growth Factors, and Proteases in Acute and Chronic Wounds. Wound Repair Regen. 1996, 4, 411–420. [Google Scholar] [CrossRef]
  285. Soehnlein, O.; Steffens, S.; Hidalgo, A.; Weber, C. Neutrophils as Protagonists and Targets in Chronic Inflammation. Nat. Rev. Immunol. 2017, 17, 248–261. [Google Scholar] [CrossRef] [PubMed]
  286. Thamm, O.C.; Koenen, P.; Bader, N.; Schneider, A.; Wutzler, S.; Neugebauer, E.A.; Spanholtz, T.A. Acute and Chronic Wound Fluids Influence Keratinocyte Function Differently. Int. Wound J. 2015, 12, 143–149. [Google Scholar] [CrossRef] [PubMed]
  287. Stojadinovic, O.; Pastar, I.; Vukelic, S.; Mahoney, M.G.; Brennan, D.; Krzyzanowska, A.; Golinko, M.; Brem, H.; Tomic-Canic, M. Deregulation of Keratinocyte Differentiation and Activation: A Hallmark of Venous Ulcers. J. Cell Mol. Med. 2008, 12, 2675–2690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  288. Bucalo, B.; Eaglstein, W.H.; Falanga, V. Inhibition of Cell Proliferation by Chronic Wound Fluid. Wound Repair Regen. 1993, 1, 181–186. [Google Scholar] [CrossRef]
  289. Cha, J.; Kwak, T.; Butmarc, J.; Kim, T.-A.; Yufit, T.; Carson, P.; Kim, S.-J.; Falanga, V. Fibroblasts from Non-Healing Human Chronic Wounds Show Decreased Expression of Βig-H3, a TGF-β Inducible Protein. J. Dermatol Sci. 2008, 50, 15–23. [Google Scholar] [CrossRef]
  290. Eming, S.A.; Krieg, T.; Davidson, J.M. Inflammation in Wound Repair: Molecular and Cellular Mechanisms. J. Investig. Dermatol. 2007, 127, 514–525. [Google Scholar] [CrossRef] [Green Version]
  291. Yu, R.; Zhang, H.; Guo, B. Conductive Biomaterials as Bioactive Wound Dressing for Wound Healing and Skin Tissue Engineering. Nano-Micro Lett. 2021, 14, 1. [Google Scholar] [CrossRef]
  292. Olingy, C.E.; San Emeterio, C.L.; Ogle, M.E.; Krieger, J.R.; Bruce, A.C.; Pfau, D.D.; Jordan, B.T.; Peirce, S.M.; Botchwey, E.A. Non-Classical Monocytes Are Biased Progenitors of Wound Healing Macrophages during Soft Tissue Injury. Sci. Rep. 2017, 7, 447. [Google Scholar] [CrossRef] [Green Version]
  293. Dario, M.D.; Colombo, E.; Govi, C.; Feo, D.D.; Messina, M.J.; Romeo, M.; Sangalli, F.; Moiola, L.; Rodegher, M.; Martino, G.; et al. Myeloid Cells as Target of Fingolimod Action in Multiple Sclerosis. Neurol.-Neuroimmunol. Neuroinflamm. 2015, 2, e157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  294. Lim, N.S.J.; Sham, A.; Chee, S.M.L.; Chan, C.; Raghunath, M. Combination of Ciclopirox Olamine and Sphingosine-1-Phosphate as Granulation Enhancer in Diabetic Wounds. Wound Repair Regen. 2016, 24, 795–809. [Google Scholar] [CrossRef] [PubMed]
  295. Awojoodu, A.O.; Ogle, M.E.; Sefcik, L.S.; Bowers, D.T.; Martin, K.; Brayman, K.L.; Lynch, K.R.; Peirce-Cottler, S.M.; Botchwey, E. Sphingosine 1-Phosphate Receptor 3 Regulates Recruitment of Anti-Inflammatory Monocytes to Microvessels during Implant Arteriogenesis. Proc. Natl. Acad. Sci. USA 2013, 110, 13785–13790. [Google Scholar] [CrossRef] [Green Version]
  296. Sun, G. Pro-Regenerative Hydrogel Restores Scarless Skin during Cutaneous Wound Healing. Adv. Healthc. Mater. 2017, 6, 1700659. [Google Scholar] [CrossRef] [PubMed]
  297. Waters, M.; VandeVord, P.; Van Dyke, M. Keratin Biomaterials Augment Anti-Inflammatory Macrophage Phenotype in Vitro. Acta Biomater. 2018, 66, 213–223. [Google Scholar] [CrossRef]
  298. Tong, W.Y.; bin Abdullah, A.Y.K.; binti Rozman, N.A.S.; bin Wahid, M.I.A.; Hossain, M.S.; Ring, L.C.; Lazim, Y.; Tan, W.-N. Antimicrobial Wound Dressing Film Utilizing Cellulose Nanocrystal as Drug Delivery System for Curcumin. Cellulose 2018, 25, 631–638. [Google Scholar] [CrossRef]
  299. Berce, C.; Muresan, M.-S.; Soritau, O.; Petrushev, B.; Tefas, L.; Rigo, I.; Ungureanu, G.; Catoi, C.; Irimie, A.; Tomuleasa, C. Cutaneous Wound Healing Using Polymeric Surgical Dressings Based on Chitosan, Sodium Hyaluronate and Resveratrol. A Preclinical Experimental Study. Colloids Surf. B Biointerfaces 2018, 163, 155–166. [Google Scholar] [CrossRef]
  300. Kasiewicz, L.N.; Whitehead, K.A. Silencing TNFα with Lipidoid Nanoparticles Downregulates Both TNFα and MCP-1 in an in Vitro Co-Culture Model of Diabetic Foot Ulcers. Acta Biomater. 2016, 32, 120–128. [Google Scholar] [CrossRef]
  301. Sridharan, R.; Cameron, A.R.; Kelly, D.J.; Kearney, C.J.; O’Brien, F.J. Biomaterial Based Modulation of Macrophage Polarization: A Review and Suggested Design Principles. Mater. Today 2015, 18, 313–325. [Google Scholar] [CrossRef]
  302. Kim, J.E.; Lee, J.H.; Kim, S.H.; Jung, Y. Skin Regeneration with Self-Assembled Peptide Hydrogels Conjugated with Substance P in a Diabetic Rat Model. Tissue Eng. Part A 2018, 24, 21–33. [Google Scholar] [CrossRef]
Figure 1. Induction and resolution of acute inflammation during tissue repair. Pathogen/Damage Associated Molecular Patterns (PAMPs/DAMPs) associated with an injury bind to Toll-like receptors (TLRs) expressed by tissue-resident cells, including macrophages. These cells release proinflammatory mediators and chemokines to activate an acute inflammatory program, recruiting leukocytes from nearby blood vessels. Neutrophils and monocytes gradually infiltrate the injury site to exert antimicrobial mechanisms, including increased reactive oxygen species (ROS) production. Following the eradication of pathogens, neutrophils undergo apoptosis to be engulfed by macrophages (efferocytosis). Activated macrophages endure polarization to release anti-inflammatory cytokines leading to resolution of inflammation and reduction in ROS levels. Moreover, a retrograde migration of neutrophils aids in decreasing infiltrating leukocytes. Additionally, macrophages release various growth factors to trigger tissue repair machinery.
Figure 1. Induction and resolution of acute inflammation during tissue repair. Pathogen/Damage Associated Molecular Patterns (PAMPs/DAMPs) associated with an injury bind to Toll-like receptors (TLRs) expressed by tissue-resident cells, including macrophages. These cells release proinflammatory mediators and chemokines to activate an acute inflammatory program, recruiting leukocytes from nearby blood vessels. Neutrophils and monocytes gradually infiltrate the injury site to exert antimicrobial mechanisms, including increased reactive oxygen species (ROS) production. Following the eradication of pathogens, neutrophils undergo apoptosis to be engulfed by macrophages (efferocytosis). Activated macrophages endure polarization to release anti-inflammatory cytokines leading to resolution of inflammation and reduction in ROS levels. Moreover, a retrograde migration of neutrophils aids in decreasing infiltrating leukocytes. Additionally, macrophages release various growth factors to trigger tissue repair machinery.
Ijms 24 00641 g001
Figure 2. A classic wound healing model shows the importance of transitioning from the inflammatory to proliferative phases during the repair process. Normal wound healing typically progresses via activation of an innate immune program involving inflammatory cell recruitment. Anti-inflammatory responses are triggered through several mechanisms, including efferocytosis, to suppress inflammation and initiate repair events. These reparative phases comprise re-epithelialization, granulation tissue formation, angiogenesis and collagen deposition. Dysregulation of inflammatory responses in case of, for example, diabetes, aging and immunosuppressive diseases fail to transition to the proliferation phase, thereby inducing delayed wound healing or chronic injuries.
Figure 2. A classic wound healing model shows the importance of transitioning from the inflammatory to proliferative phases during the repair process. Normal wound healing typically progresses via activation of an innate immune program involving inflammatory cell recruitment. Anti-inflammatory responses are triggered through several mechanisms, including efferocytosis, to suppress inflammation and initiate repair events. These reparative phases comprise re-epithelialization, granulation tissue formation, angiogenesis and collagen deposition. Dysregulation of inflammatory responses in case of, for example, diabetes, aging and immunosuppressive diseases fail to transition to the proliferation phase, thereby inducing delayed wound healing or chronic injuries.
Ijms 24 00641 g002
Table 2. Functions of macrophage phenotypes during tissue repair.
Table 2. Functions of macrophage phenotypes during tissue repair.
PhenotypeReceptorsFunctions
M1 (classically activated or proinflammatory)CD68
CD86 CD80
Induces microbicidal activities (NO, ROS and phagocytosis) [187].
Releases proinflammatory cytokines (TNF-α, IL-1β and IL-6) [178,193]
Enhances neutrophil recruitment by expressing chemokines [193,194] and synthesizing MMPs to degrade ECM [195]
Clears apoptotic and necrotic PMN [196]
M2a (alternatively activated or wound healing)CD163
CD206 CD209Ym1
Activated by IL-4/IL-13 [197]
Produces chemokines: CCL17, CCL18 and CCL22 as well as growth factors: IGF-1, fibronectin, TGF-β and PDGF [198,199]
Promotes ECM formation and angiogenesis [200]
M2b (regulatory or type 2)CD86
Activated in vitro by phagocytosing apoptotic neutrophils [201]
Inhibits inflammation by releasing IL-10 [201]
Expresses IL-6, CCL1 and high levels of iNOS [198,202]
M2c (pro-resolving or deactivated)CD86
CD163
CD206
Stimulated by IL-10 via STAT3 pathway
Releases IL-10 and TGF-β to exhibit anti-inflammatory responses [9,203]
Expresses Mer receptor tyrosine kinase (MerTK), essential for efferocytosis [204]
M2d (tumor-associated macrophages)-
Activated by IL-6 or both TLR ligands and A2 adenosine receptor agonists [205,206]
Secretes high levels of VEGF, IL-10 and TGF-β and downregulates TNF-α, IL-12 and IL-1β [207,208]
TNF-α: tumor necrosis factor-alpha; PMN: polymorphonuclear leukocytes; IL: interleukin; TGF-β: transforming growth factor-beta; NO: nitric oxide; ROS: reactive oxygen species; MMPs: matrix metalloproteinases; ECM: extracellular matrix; CCL: chemokine (C-C motif) ligand; IGF-1: insulin growth factor-1; PDGF: platelet-derived growth factor; iNOS: inducible nitric oxide synthase; STAT3: signal transducer and activator of transcription 3; TLR: Toll-like receptors; VEGF: vascular endothelial growth factor.
Table 3. Contributions of mast cells to tissue repair.
Table 3. Contributions of mast cells to tissue repair.
Repair EventFunctions
Hemostasis
  • Release tryptase to deactivate clotting induced by thrombin-stimulated fibrinogen; while they can also express fibrin stabilizing factor (XIIIa) to strengthen cross-linking of fibrin fibrils [228]
  • Produce plasminogen activator inhibitor 1, a potent inhibitor of fibrinolysis, upon their activation by complements (C5a) [238]
  • Enhance vascular permeability via heparin-mediated increase in bradykinin [239]
Inflammation
  • Induce vasodilation and increase vascular permeability to enhance cellular influx [240,241]
  • Secrete MCP-1 to recruit monocytes [16,242] as well as TNF-α, CXCL8, kinins and proteases to recruit neutrophils [241,243,244]
  • Promote the release of human β-defensin-3, an antimicrobial peptide, from epidermal keratinocytes [245]
  • MCs-secreted histamine increases PPRs (TLR-2 and Dectin-1) expression on keratinocytes and raises levels of GM-CSF and CXCL8 [245]
Re-epithelialization
  • MCs-derived tryptase facilitates interactions between keratinocytes and MCs via degrading ECM and triggering PAR-2 receptors on keratinocytes [246]
  • MCs-released LTB4 enhance keratinocyte proliferation [247], while MCs-produced histamine inhibits it [248]
Granulation tissue formation
  • Promote fibroblast proliferation through secreting bFGF, IL-4 and VEGF [229,249]
  • MCs-derived histamine and tryptase stimulate fibroblasts to release FGF-2 or FGF-7 [229]
Angiogenesis
  • MCs-derived tryptase degrades the basement membrane to allow for endothelial cells migration and proliferation [250]
  • Secret a variety of pro-angiogenic mediators such as PDGF, VEGF, FGF-2, bFGF, ANG-1 and TGF-β [236,250]
TNF-α: tumor necrosis factor-alpha; CXCL8: C-X-C Motif Chemokine Ligand 8; PPRs: pattern recognition receptors; TLR: roll-like receptor; PAR2: protease-activated receptor 2; LTB4: leukotriene B4; bFGF: basic fibroblast growth factor; VEGF: vascular endothelial growth factor; PDGF: platelet-derived growth factor; ANG-1: angiopoietin-1.
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

Soliman, A.M.; Barreda, D.R. Acute Inflammation in Tissue Healing. Int. J. Mol. Sci. 2023, 24, 641. https://doi.org/10.3390/ijms24010641

AMA Style

Soliman AM, Barreda DR. Acute Inflammation in Tissue Healing. International Journal of Molecular Sciences. 2023; 24(1):641. https://doi.org/10.3390/ijms24010641

Chicago/Turabian Style

Soliman, Amro M., and Daniel R. Barreda. 2023. "Acute Inflammation in Tissue Healing" International Journal of Molecular Sciences 24, no. 1: 641. https://doi.org/10.3390/ijms24010641

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