Next Article in Journal
Th17-Dependent Nasal Hyperresponsiveness Is Mitigated by Steroid Treatment
Next Article in Special Issue
Targeting Nrf2 with Probiotics and Postbiotics in the Treatment of Periodontitis
Previous Article in Journal
CAST as a Potential Oncogene, Identified by Machine Search, in Gastric Cancer Infiltrated with Macrophages and Associated with Lgr5
Previous Article in Special Issue
p62/SQSTM1/Keap1/NRF2 Axis Reduces Cancer Cells Death-Sensitivity in Response to Zn(II)–Curcumin Complex
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Loricrin at the Boundary between Inside and Outside

1
Department of Dermatology Integrated Medicine, Osaka University Graduate School of Medicine, 2-2 Yamadaoka, Suita 565-0871, Osaka, Japan
2
Department of Dermatology and Charles C. Gates Center for Regenerative Medicine, University of Colorado Anschutz Medical Campus, Aurora, CO 80045, USA
*
Author to whom correspondence should be addressed.
Biomolecules 2022, 12(5), 673; https://doi.org/10.3390/biom12050673
Submission received: 13 April 2022 / Revised: 30 April 2022 / Accepted: 4 May 2022 / Published: 6 May 2022

Abstract

:
Cornification is a specialized mode of the cell-death program exclusively allowed for terrestrial amniotes. Recent investigations suggest that loricrin (LOR) is an important cornification effector. As the connotation of its name (“lorica” meaning an armor in Latin) suggests, the keratin-associated protein LOR promotes the maturation of the epidermal structure through organizing covalent cross-linkages, endowing the epidermis with the protection against oxidative injuries. By reviewing cornification mechanisms, we seek to classify ichthyosiform dermatoses based on their function, rather than clinical manifestations. We also reviewed recent mechanistic insights into the Kelch-like erythroid cell-derived protein with the cap “n” collar homology-associated protein 1/nuclear factor erythroid 2-related factor 2 (NRF2) signaling pathway in skin health and diseases, as LOR and NRF2 coordinate the epidermis-intrinsic xenobiotic metabolism. Finally, we refine the theoretical framework of cross-talking between keratinocytes and epidermal resident leukocytes, dissecting an LOR immunomodulatory function.

1. Introduction and Overview

1.1. Cornification: A Successful Tissue Adaptation to Land

The body’s surface is covered with a specialized protective layer called the epidermis, from the scalp to the foot: the stratified squamous epithelium consisting of keratinocytes (KCs). Mammalian epidermises acquire specialized apparatuses to accommodate various external stresses, including physical injuries, harmful substances, or ionizing radiation. The stratum corneum (SC) overlies most parts of the epidermis (interfollicular epidermis (IFE)) and protects against dehydration or mechanical loads. Interspersed in the epidermis are skin appendages that also produce cornified materials. The hair provides the integument with insulation against heat/cold or trauma in mammals, and comprises the hairy (fibrillar) attachment system in the reptilian foot [1]. The mammalian nail comprises essential components of the haptile, tactile, mechanic, and the defensive and/or predatory instrumentarium [2].
These end-products are derived from the functional cell-death program termed cornification [3]. The IFE and appendages serve different functionalities by producing the SC (armor) and the hair/nails (instruments), respectively. The effector proteins (keratins and keratin-associated proteins) exhibit various characteristics to meet the biophysical/biochemical properties required for specific body parts. Despite the substantial difference in appearance, they are the macromolecular “protein alloy”, comprising the extensive intra/intermolecular formation of ε-(γ-glutamyl) lysine isopeptide [4,5]/disulfide (–S–S–) cross-linkages [6,7]. Noteworthily, these vital biochemical characteristics are observed even in the simple columnar epithelium, covering most parts of the wet-surfaced digestive tract. It was recently proven that gel-forming mucins (Mucin 2, MUC2), secreted from the goblet cells, undergo transglutaminase 3 (TGM3)-mediated isopeptide cross-linkages [8], which are resistant to reducing reagents that cleave disulfide (–S–S–) cross-linkages [9]. The results are a reminder of the very early observations that the cell membrane-like remnants of the “mummified” KC (corneocyte), i.e., cornified cell envelopes (CEs), are highly resistant to exhaustive hydrolysis [10] because of their extremely strong isopeptide cross-linkages [5]. Therefore, surface protection by turning secreted substances into macromolecules may be a strategy conserved across tissues.
Analogously to the ontogenetic evidence, vertebrates have acquired gene sets that encode the effector proteins more suitable for cornification than the pre-existing repertoire through gene duplication or deletion during evolution [11,12]. For instance, when amphibians invaded arboreal habitats on the waterside [13], ancestral keratins are thought to have emerged as constituents of the toe pad, which endure the forces caused by the repetitive contact with the waterside environment [13,14,15]. Thus, the ancient keratins are considered to have evolved into a major group of cysteine-rich alpha keratins, enabling the hairy (fibrillar) attachment system in reptiles [1] and constituting mammalian hair [13]. Despite the presence of amphibian CEs [14] or cornified layers [16] in the epidermis, phylogenetic evidence strongly supports the notion that cornification is an ultimate form of body wall protection that is exclusively allowed for terrestrial amniotes (reptiles, birds, and mammals) [17,18,19,20,21]. A major cornified cell envelope protein, loricrin (LOR), is considered to have originated in a common ancestor of modern amniotes, along with the sauropsid-specific corneous beta-proteins (beta keratins) that constitute the reptilian/avian SC [12,21]. Aggregately, the major product of cornification SC is arguably an example of successful tissue adaptation to land.

1.2. Cornification in Skin Health and Diseases

Several diseased conditions of inflammatory or non-inflammatory causes comprise aberrations in cornification. Psoriasis, eczema (atopic dermatitis (AD)), or acne may exemplify the former, and calluses, corns, or ichthyoses may represent the latter. As a higher-order structure with dynamics, the epidermal tissue produces pro-inflammatory cytokines in response to primary injuries, and direct pathogen recognition via the toll-interleukin receptor (TIR) [22]. This frontline protective immunity can provide critical directions for such diseases. Indeed, therapeutic intervention in the tissue response cascades has been successful for AD and psoriasis by targeting the interleukin (IL)-4/IL-13 receptor and the IL-23/IL-17 signaling pathway, respectively [23]. Omic approaches, such as single-cell technology, have provided us with high-resolution images of cutaneous pathologies [24]. Data-mining approaches may boost the identification of potential therapeutic targets and depict the evolutionary trajectory of various diseased tissues, including skin cancers [25] or inflammatory conditions [26]. Dissecting the cause–effect relationships between the diseases and the genetic predisposition factors, including the loss of structural proteins (e.g., filaggrin (FLG)-deficiency (ichthyosis vulgaris—IV) in AD [27]) or unchecked TIR signaling cascades (e.g., hyperactivated caspase recruitment domain family member 14 (CARD14) in psoriasis [28]) may remind us of the premise that most ailments are products of gene–environment mismatches [29]. In this article, we focused on LOR and NFE2-related factor 2 (NRF2), which are important cornification effectors in IFE, and sought to revise the theoretical framework of the cross-talk between IFE KCs and leukocytes proposed previously [30,31,32,33].

2. SC Permeability Barrier: The Priority

2.1. IFE Cornification in Brief

In the homeostatic status, lineage-committed IFE KCs [25] exit from the proliferative layer (stratum basale—SB), detach from the basement membrane [34], and migrate upward through the stratum spinosum and the stratum granulosum (SG). SG is morphologically identified by the presence of keratohyalin granules, and contains keratin-associated proteins FLG and LOR [35]. Tight junctions (TJs) connect the plasma membrane of adjacent KCs and form a belt-like structure (zonula) [36], denoting the apical end of the living epidermal layers. The SG comprises three distinct layers arranged in an apicobasal direction. The uppermost SG1 is particularly important because IFE KCs undergo the specialized cell-death program cornification (or “corneoptosis” [37]). The specialized body wall defense program involves a myriad of biological processes. These include the transient influx of calcium (Ca2+) [38], intracellular acidification [37], autophagy/nucleophagy [39], activated proteases [40,41,42,43,44], vitamin A (VA) metabolism [45]/vitamin D metabolism [46], the lamellar granule (LG)-mediated secretion of intracellular content (lipids/proteins [47,48]) to the interstitial space [49], the fusion of LG-limiting membranes with apical plasma membranes [50,51], TJ disappearance [36], dephosphorylation of the huge pro-FLG protein and the subsequent protease-mediated breakdown into FLG monomers (reviewed in ref [52]), TGM-mediated cytoskeletal cross-linkages [53,54] and disulfide (–S–S–) cross-linkages [55,56,57,58,59,60,61,62], etc.
After leaving the transitional layer, IFE KCs turn into a polymerized epithelial sheet composed of “mummified” KCs (corneocytes). Corneocytes are surrounded by orderly arranged lipids (ceramides) covalently attached to the outside of CEs (corneocyte lipid envelopes (CLE)), and the structure of the SC is analogized to “bricks and mortar” [63]. Despite being “dead”, the lower layer (stratum compactum) appears important, since corneocytes continue structural maturation before shedding off the skin surface. This process involves cross-linkages of cytoskeletal and junctional components (desmosomal-keratin scaffold (DKS) of the SC) with CEs. The former includes differentiated keratins K1/K10 and FLG monomers, and the latter corresponds to corneodesosin (CDSN). The major CE protein, LOR, moves to the desmosomal area at the periphery [64,65], reinforces the DKS [66] via the formation of intermolecular disulfide (–S–S–) cross-linkages, promotes corneocyte compactization/flattening, and transitions into the stratum disjunctum, ready for peeling away [56,57,67].

2.2. FLG: A Precursor of Natural Moisturizing Factors

The relatively immature stratum compactum appears to absorb small water-soluble molecules with the help of humectants (so-called natural moisturizing factors (NMFs) [52]) stored in corneocytes. Because the FLG monomers are major precursors of hygroscopic (moisture-absorbing) amino acids, such as arginine (Arg) [52], they possess a much smaller amount of Arg in the stratum compactum and fail to confer impermeability against small water-soluble molecules that could be absorbed by corneocytes, although FLG-knockout mice exhibited an abnormal aggregation of keratins, as the name connotes (i.e., filament aggregation [67]). Metal ions are the established water-soluble substances that cause allergic contact dermatitis (ACD). Consequently, FLG-deficient mouse models exhibited hypersensitive reactions [68,69] similar to IV patients [70], and were susceptible to Staphylococcus aureus (S. aureus) colonization [71]. The stratum compactum may correspond to the aseptic mucous layers composed of MUC2 polymers that prevent the penetration of commensal bacteria and subsequent T helper (TH) 17-mediated immunopathology that models inflammatory bowel diseases [8].

2.3. Corneodesmosomes: The Regulator of Corneocyte Cohesion

The SC is a polymerized sheet consisting mainly of keratins and keratin-associated proteins. In addition to cross-linkage and the subsequent breakdown of FLG monomers, desmosomal components must be degraded quickly [72,73]. Serine proteases, SC chymotryptic enzymes (kallikrein 7—KLK7), and SC tryptic enzymes (SCTE; KLK5) target desmosomal components [74]. A serine protease inhibitor, lymphoepithelial Kazal-type related inhibitor type 5 (LEKTI), antagonizes KLK activity and prevents the premature proteolysis of the N-terminal extracellular domain of desmoglein 1 (DSG1)/CDSN. Unopposed desmosomal breakdown leads to postnatal lethality in mice [40,43,75,76,77,78] and severe allergic manifestations that include AD-like skin phenotypes in humans [79,80]. Likewise, the loss-of-function (LOF) variants in DSG1/CDSN phenocopy the allergic symptoms in humans (reviewed in [81] and discussed later). Although a defective TJ-based permeability barrier can cause ichthyosiform skin phenotypes in humans [82], this aggregated evidence suggests that the DKS of the SC [66] assures successful cornification.

2.4. A classification of Ichthyosiform Dermatoses Based on Gene Functions

The aggregated molecular genetic evidence regarding rare, autosomal recessive ichthyosiform dermatoses suggests the importance of “the mortar” for preventing dehydration. We took autosomal recessive congenital ichthyosis (ARCI), peeling skin syndrome-1 (PSS1) (Online Mendelian Inheritance in Man (OMIM) #270300, accessed on 1 April 2022 [83]), Netherton syndrome (NS; OMIM #256500) [79], skin dermatitis, multiple severe allergies, and metabolic wasting (SAM) syndrome (OMIM #615508) [80] as clinical examples. Although Vohwinkel syndrome (VS) with ichthyosis (OMIM #604117) [84] is not an autosomal recessive trait, we took the rare autosomal dominant dermatosis to clarify the LOR effector function in cornification.

2.4.1. ARCI

ARCIs are heterogeneous groups of rare and congenital ichthyoses typically characterized by a parchment-like collodion membrane at birth and subsequent widespread hyperkeratosis covering almost the entire body surface (reviewed in [85]). Hyperkeratosis severity may vary depending on the afflicted individual, ranging from skin flakiness (congenital ichthyosiform erythroderma, CIE) to thick, large, and plate-like squames (lamellar/harlequin ichthyoses, LI/HI). Although there are no established classifications based on gene functions, defects in the following biological processes have been suggested: i. lipid metabolism (ceramide synthesis) [86,87,88,89,90,91,92,93]; ii. LG function [94]; iii. CLE formation [95]; and iv. FLG processing [96,97] (Table 1).

2.4.2. Desmosomal Defects Accompanying Allergic Manifestations

NS [79] and PSS1 [98] are similar in that CDSN deficiencies cause both conditions, by unopposed degradation and absence, respectively. LOF variants in the DSG1 gene cause SAM syndrome [80] (Table 2).
The molecular pathology of these skin-peeling disorders provides a profound insight into AD pathogenesis [81]. NS is clinically characterized by ichthyosiform erythroderma, hair shaft abnormality, and atopic manifestations. The patients are born with erythroderma that develops into ichthyosis linearis circumflexa (ILC) [99]. Systemic complications are also common, including failure to thrive (FTT), eosinophilia, increased serum IgE levels, systemic infection, and food allergies [81]. LOF variants cause NS in the serine protease inhibitor of the Kasal-type 5 (Spink5) gene that encodes LEKTI. Unchecked proteolytic activity of KLKs promotes CDSN/DSG1 extracellular domain degradation [81]. Recapitulating the lethal phenotype results in SC detachment [43], and forced KLK5 overexpression in the epidermis faithfully replicated major NS symptoms in mice [40]. LOF variants cause PSS in the CDSN, clinically characterized by ichthyosiform erythroderma, FTT, increased IgE levels, eosinophilia, asthma, urticaria/angioedema, and food allergies [81]. Currently, two independent gene knockout studies of CDSN have been reported [77,78], both resulting in postnatal lethality that phenocopies the Spink5-knockout mice. LOF variants cause SAM syndrome in the DSG1 gene, clinically characterized by striate palmoplantar keratoderma (PKK), FTT, and atopic manifestations such as NS or PSS1. SAM syndrome-associated PKK is a distinctive clinical feature among these desmosomal disorders, suggesting that DSG1 serves as an indispensable component of the DKS in the differentiating epidermal layer, which, along with the K1/K10 pair [100], resists shear stress in the weight-bearing skin, i.e., the palm and sole [101]. DSG1-knockout mice exhibit postnatal lethality due to a superficial detachment of the epidermal tissue [75,76], phenocopying the NS/PSS1 mouse models [40,43,78,79].

2.4.3. Ichthyosiform Dermatosis Caused by LOR Mislocalization

VS with ichthyosis is an autosomal dominant ichthyosiform dermatosis caused by LOR gene variants [84]. Characteristics include palmoplantar hyperkeratosis with small “honeycomb” depressions and the progressive formation of pseudoainhum (digital constricting bands) [84] or erythrokeratodermia [102]. VS-related LOR mutations result in delayed translation termination, and the C-terminal glycine- or glutamine/lysine-rich domain is replaced with arginine and leucine-rich amino acid sequences. In contrast to wild-type (WT) LOR that moves to the cell periphery and forms CE [35], mutant LOR accumulates in the nucleus [102]. However, LOR glutamine/lysine residues required for TGM-mediated isopeptide cross-linkages should not be affected by the VS mutation located at the C-terminus [60]. Indeed, the LOR C-terminal mutation generates a nuclear localization signal, and abnormal nuclear accumulation of the LOR perturbs normal cornification, causing an impaired SC permeability barrier assessed by transepidermal water loss or the passive diffusion of water-soluble Lucifer yellow fluorescent dye [103]. We have previously demonstrated the dominant negative effect of the vs. mutation by generating the VS-Lor-transgenic mice with an LOR-knockout (LKO) background, which phenocopied VS-Lor-transgenic mice with WT LOR [103] (Table 3).

2.4.4. Potential Targeted Therapeutics Based on the Gene Expression Profiles

Despite clear cause–effect relationships proven by molecular genetic studies, treating ichthyosiform dermatoses remains unsatisfactory [104]. Most conditions are treated with supportive regimens (skin moisturizers or emollients) or VA derivatives [104]. As in the proposal of “psoriasiform ichthyosis”, referring to NS/ILC [105], extensive gene expression profiling using relatively small skin biopsy or blood samples has uncovered TH17-skewing as an immunological signature shared across ichthyosiform dermatoses [104,106].
The IL-23/IL-17 axis executes body surface pathogen clearance through the prompt recruitment of neutrophils [107], with psoriasis (psoriasiform tissue reactions [108]) being an epitome of such tissue responses. Therefore, one might have expected the results from the omics approach, given that defects in the epidermal structure cause most ichthyosiform dermatoses and can activate the primary immune defense of the skin (so-called “preemptive” immunity [109]) imprinted within the tissue [110,111,112]. Although DSG1-knockout mice exhibit postnatal lethality due to a superficial detachment of the epidermal tissue [75,76], they exhibit TH17-associated inflammatory signatures on embryonic day 18.5 (right before birth) [75]. Thus, results from the rodent models and SAM syndrome patients may correspond to immunological alertness [109] augmented by the breach of superficial epidermal tissue [71]. In fact, therapeutic interventions on the IL-23/IL-17 axis in NS [113,114,115,116] or SAM syndrome [75,115] patients have shown success in clinical anecdotes. Based on a recent small-size pilot study, Guttman-Yassky and Paller et al. concluded that younger and more erythrodermic patients will benefit from these immunotherapies [115]. Future studies are needed to clarify endotypic or phenotypic differences in responders vs. non-responders. Clinical studies will increase our understanding of disease heterogeneity and lead to the precise prediction of therapeutic responses [115].

3. LOR as a Major Epidermal Differentiation Component

LOR and NRF2 in IFE Cornification

Because LOR is a major CE protein appearing from the cytosol at the later phase of cornification and replacing the plasma membrane [50], the largely asymptomatic LKO phenotype might not have been surprising retrospectively [59]. We previously found that LKO mice exhibit a delay in the maturation of the epidermal permeability barrier assessed by Toluidine blue dye exclusion in utero [59]. The LKO embryo showed a massive upregulation of small proline-rich proteins 2 (SPRR2s) [58,59]. The Sprr/SPRR genes encoding the cytoprotective proteins reside within the epidermal differentiation complex (EDC) [117] located in the mouse chromosome 3/human chromosome 1, clustering with other cornification proteins, including LOR, involucrin, and FLG [30]. SPRR proteins are rich in thiol (–SH), show structural resemblance to LOR [118], and are recently shown to have antimicrobial properties in extracutaneous tissues, including the gut [119] and the kidney [120]. Owing to the abundance of thiol (–SH) residues, SPRR2s possess superior antioxidative capacity compared to LOR in vitro [121], which may help vertebrates recover from physical insults, including wounding [122], ultraviolet (UV) [123], or the invasion of metazoan parasites [119]. Along the same line, late-cornified cell envelope protein 1s (LCE1s) are also abundant in LKO CE [118]. Recently, it was proven in humans that LCE3 possess a β-defensin-like broad antimicrobial activity in vitro and, like SPRRs, they are massively induced following superficial epidermal injury [124] or psoriatic epidermis [125,126].
The Kelch-like erythroid cell-derived protein with the cap “n” collar homology-associated protein 1 (KEAP1)/NFE2-related factor 2 (NRF2) signaling pathway is a ubiquitous antioxidative machinery that drives quick antioxidative responses following the detection of cellular oxidative insults [127]. Actin-associated cytoskeletal KEAP1 protein [128] is rich in thiols (–SH), and subjects the transcription factor NRF2 to proteasome-mediated degradation in the cytosol in the steady state [127]. Electrophilic stresses or other non-specific injuries cause conformational changes in KEAP1 proteins, stabilizing the NRF2 protein and unleashing transcriptional activity [127]. Thus, KEAP1 inhibition leads to the constitutive activation of the transcription factor, NRF2, and causes extremely strong antioxidative responses in multiple organ systems [127]. However, the Keap1-null mutation in mice leads to postnatal lethality; an autopsy of pups revealed that the upper digestive tract (the esophagus–forestomach) was obstructed with keratinous materials, and the skin manifestation somewhat resembled that of ARCI [129]. Because KEAP1/NRF2 expression levels become higher upon epidermal differentiation [130,131], the system may have evolved to convert the cellular dysregulation of various causes, including electrophiles [132,133], physical injuries [129,134], or UV [131,135], into antioxidative responses. This evidence indicates that NRF2 is a strong keratinization regulator [129,136] and reminds us of the principle that sulfur metabolism governs epidermal differentiation [6,7].
The delayed barrier formation of LKO embryo and the supposed LOR function, i.e., a major TGM substrate that promotes disulfide (–S–S–) cross-linkages in the SC [61,62] and the compensatory upregulation of the antioxidative SPRR2s, prompted us to test the role of the KEAP1/NRF2 signaling pathway in the LKO epidermis [55,58,118,137]. Because LOR-NRF2 double-knockout (DKO) mice look apparently healthy [55], we introduced the dominant-negative Nrf2Nrf2) gene under the control of the Lor promoter to obtain differentiating layer-specific gene expression [58]. Compared with normal-looking WT-ΔNrf2 mice, LKO-ΔNrf2 mice exhibited postnatal lethal dehydration [58]. Correspondingly, we recently found that the DKO mice exhibit a significant delay in the recovery of the SC permeability barrier following SC removal [55]. As was previously shown in other experimental systems in which constitutively active NRF2 was overexpressed in the epidermis [138], NRF2 causes the upregulation of the corneocyte cohesion protein CDSN [55]. Perhaps more importantly, meticulous ultrastructural observations revealed that corneodesmosomes remained undegraded in LKO/DKO mice compared to WT/Nrf2-knockout mice, while CEs were thinner in LKO mice than in WT mice [55]. These results may correspond to the supposed LOR kinetics in vivo: i. moving to the cell periphery and desmosomal area [64,65]; ii. forming CE and promoting corneocyte maturation [55,56,57,59,61,62]; iii. connecting LG-derived CDSN to CE via disulfide (–S–S–) cross-linkages [67,139,140]. In short, LOR appears to be the fundamental component that promotes the terminal differentiation of the epidermis [33].

4. “Structural Imprinting” of the Cutaneous Immune Effector Functions

4.1. Breach of SC Permeability Drives Inflammation

We have stressed the importance of the SC permeability barrier for the defects in cornification. ARCI is caused by defects in the following: (i) lipid metabolism (ceramide synthesis); (ii) LG function; (iii) CLE formation; and (iv) FLG processing, while desmosomal defects cause allergic manifestations [81]. Regarding LOR function, we have characterized that LOR is dispensable for the SC permeability barrier per se, unless it is mislocalized in the nucleus [103]. Nonetheless, LKO CEs are substantially thinner than WT [55], abundant with the “scaffolding” components that form CLEs [57] and have incomplete K1/K10 cross-linkages ([57] and unpublished observations). Although LKO CE phenotypes faithfully mirror Steinert’s model of CE assembly [50], investigations regarding the mechanism that regulates the desmosomal disappearance or skin peeling (desquamation) were in a nascent stage [141]. Moreover, CE isolation from the human skin hampered the analysis of the epidermal differentiation dynamics, involving the cellular turnover. Therefore, retrospectively, the elegant model suffered from a paucity of information regarding how desmosomal components are broken down [50,142] (although desmosomal components are the initial “scaffolding” component [143]). Given that TGM1-knockout mice do not generate CEs due to impaired LOR polymerization at the cell periphery [54,139], it would have been reasonable that the largely asymptomatic phenotype observed in LKO mice was surprising [140]. A more recent characterization of lethal Cdsn [77,78]/Dsg-1 [75,76]-knockout phenotypes suggests that corneodesmosomes are indispensable for successful cornification. An intriguing question arises here: why do defects in corneodesmosomal components cause allergic inflammation that resembles AD [81]? A recent Dsg1-knockout study suggests that this immunological skewing emerges as early as on embryonic day 18.5, when the epidermis fully matures, starts to turn over (degradation of the SC components: FLG or CDSN/DSG1), and becomes ready for the terrestrial life [144,145].

4.2. Epidermis-Directed Immune Responses following Breached Barrier Function

It is known that the epidermis is a source of danger signals, causing immunological alertness and subsequent tissue responses [146]. For instance, the microbiological milieu of epidermal inclusion cysts does not differ whether inflamed or not [147], presumably because of the SC’s properties evoking strong inflammatory responses. The SC causes neutrophilic abscess when subcutaneously implanted [148], and its aqueous extract causes pyrexia when administered intravenously [149] in mice. This strong SC pro-inflammatory property appears attributable to IL-1α, which is shown to be abundant in individuals with ichthyosis vulgaris (caused by LOF FLG variants) [150], or LI [151], in close association with the maturation/recovery of the epidermal barrier [152,153]. Although IL-1α and IL-1β are prototypes of leaderless cytokines lacking signal peptides for secretion, IL-1α does not require an inflammasome for activation [154]. The pro-IL-1α stored in the nucleus is also biologically active [155] and is released in response to physical injuries such as mechanical deformation [156]. Therefore, IL-1α, abundantly present in the SC, serves as a damage-associated molecular pattern, initiating ACD [157] or AD-like inflammation in the “flaky tail” mice [158]. It appears most likely that IL-1α, densely present atop the epidermis, informs epidermal damage (or irritation), conditions antigen presenting cells, and enhances inflammatory responses skewed from the acute TH17 type [159] toward the chronic TH2 type [132]. Indeed, Bermekimab, a monoclonal anti-IL-1α antibody, has potential as a new therapeutic option [160], inducing rash clearance and reducing itching in patients with AD (NCT03496974). The evidence suggests that, in response to injuries, the SC generates intrinsic inflammatory cues that break tolerance and promote TH2 immunological memory, a hallmark of AD [132].

4.3. LOR: The Scaffold of Sulfur Metabolism in the IFE

Next, we must return to the principle that cornification is all about anabolizing sulfur; thiol (–SH) is effectively converted into disulfide (–S–S–) [6,7], and TGMs or their substrates constitute important effectors [54,59,143]. LOR, precipitating at the cell periphery and replacing the plasma cell membrane [35,55], confers photoprotection in corneocytes [56] and protects against polyaromatic hydrocarbon 7,12-dimethylbenz (a) anthracene (DMBA) [137]. LOR absence/presence does not interfere with CLE formation [63,161], which is preceded by the fusion of the apical plasma cell membrane with the LG-limiting membrane [51,161]. Consequently, the penetration of foreign substances via the SC paracellular route, filled with the lamellated lipids attached to CLEs, is unaltered in LKO mice [55,59,137]. In aggregate, as the connotation of its name (“lorica” meaning armor in Latin [62]), or the phylogenetic evidence [12,21] suggest, LOR constitutes an essential element of epidermal host defense, in coordination with the KEAP1/NRF2 signaling pathway [55,58,137]. The adaptor-like function also promotes the structural maturation of corneodesmosomes, SC’s specialized adhesion structure, as extensively studied and suggested previously [141,162,163,164,165].

4.4. LOR at the Boundary between Inside and Outside

The reduction and oxidation (redox) state of thiol (–SH)/disulfide (–S–S–) serves as an important indicator of the inside and outside of organisms. The cytosol is highly reduced at the cellular level, owing to a group of thiol-regulating enzymatic systems, including the thioredoxin–thioredoxin reductase and glutaredoxin–glutathione systems [166]. Conversely, the extracellular space is strongly oxidized due to the paucity of the reducing enzymes and the presence of oxygen or other gaseous mediators (such as nitric oxide) [166]. The same may be applied to the epidermal differentiation in which the redox state governs adaptive tissue responses [55,58,136,163] and homeostasis [55,56,57,118,137]. Located in the interstitial space (the paracellular route) of the differentiating epidermal layer are epidermal Langerhans cells (LCs), unique resident leukocytes with a macrophage/dendritic cell (DC) lineage (reviewed in [167,168]). LCs are prototypes of tolerogenic immature DCs [169] that express TJ proteins [170], protect against ACD [171,172], or phorbol ester-induced irritant dermatitis [173], and promote the recovery of the epidermal barrier following physical disruption [174]. The seemingly odd property for classic immune cells may be attributable to epidermal LCs’ peculiar lifestyle, which significantly differs from that of mucosal LCs [175].

4.5. Does Epidermal Differentiation Coincide with Immunological Maturation?

Earlier structural observations suggest that epidermal LCs and KCs exist at a constant ratio [176,177], constituting the epidermal proliferation unit (EPU) [176] or the epidermal differentiation unit (EDU) [34]. This strikingly constant ratio between KCs and resient leukocytes (LCs and dendritic epidermal T cells (DETCs)) [178], presumably facilitates the quality control of the differentiating KCs migrating out from the SB [179]. The epidermal residence of LC and DETC requires the transmembrane receptors αVβ6 integrin [180] and αEβ7 integrin (CD103) [181,182], respectively. Therefore, the epidermis serves as a scaffold for the transmembrane communication of these resident leukocytes, and epidermal differentiation (cornification) may efficiently terminate the physical connection. Here, we take LC as a prototype of the epidermal resident leukocyte (Figure 1).
LCs are derived from fetal hematopoietic cells and undergo homeostatic proliferation independent of the blood circulation; however, disrupted LC–KC connection by UV [180,183] evokes the repopulation of blood-borne monocyte-derived LCs to the empty niche [184,185]. Immature LCs stay in the epidermis by activating endogenous TGF-β signaling via αVβ6 integrin expressed on the plasma cell membrane of KCs [180]. Bioactive TGF-β upregulates the adhesion molecules required for epidermal retention, such as the epithelial cell adhesion molecule (EpiCAM) [186,187] and E-cadherin (ECAD) [188]. Conversely, the disrupted integrin-mediated anchorage, not prototypic inflammatory stimuli, such as TIR/nuclear factor kappa B signaling [189,190,191,192], causes LC maturation/migration, causing the transfer of antigens to the T cell-rich inner cortex of local draining lymph nodes (DLNs) [193]. This evidence, along with the fact that the monocyte–macrophage system is sensitive to extracellular redox milieu [194,195], suggests that the epidermal niche is the critical determinant of the epidermal LC phenotype [196,197]. Specifically, the LC connection with KCs [198] through the plasma cell membrane would be mechanically disrupted upon cornification, in which CEs replace the plasma cell membrane of KCs. This deduction leads to our ongoing hypothesis that LOR could profoundly affect LC phenotypes, determined by the expression levels of cell surface proteins [191,199,200], or behavioral responses following physical injury [201]. Indeed, our preliminary observations suggest that LCs in the LKO epidermis exhibit more immature but less migratory (more resident) phenotypes, resulting in the attenuation of the efferent ACD arm, a classic experimental method to determine adaptive immunity (data not shown).

5. Conclusions

Summarily, LOR, a strong effector of epidermal differentiation, directly affects LC differentiation (maturation) and thus structurally imprints cutaneous immune effector functions through TGF-β bioavailability regulation. Our hypothesis is rather simple and based on the following evidence and assumptions: 1. epidermal LCs express ECAD/TJ proteins that constitute the “apical junctional complex” and contribute to the formation of the permeability barrier with dynamics, previously proposed as the “reticuloepithelial trap” [202]. 2. The LOR replaces the plasma cell membrane following LG-limiting membrane fusion and promotes corneocyte compactization (maturation) [145]. 3. Delayed structural maturation in the LKO epidermis may retain the αVβ6-mediated LC–KC connection to the more apical end (outside layers). 4. LCs in the LKO epidermis remain mobile and better at capturing antigenic stimuli through endocytosis than LCs in the WT epidermis. 5. The immature structure of the LKO epidermis imprints adaptive immune responses through the relatively increased bioavailability of TGF-β, which may then precondition naïve CD8+ T cells [203] and suppress effector T cells during active inflammation [199,200] (Figure 1).
Further investigational studies are required. Revealing this mechanism may provide us with novel therapeutic interventions that percutaneously control immune-mediated diseases, such as multiple sclerosis [204,205] or food allergies [206,207], as an alternative to vaccination against plagues.

Author Contributions

Conceptualization, Y.I.; validation, D.R.R.; writing—original draft preparation, Y.I.; writing—review and editing, D.R.R.; supervision, D.R.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported in part by the following grants: JSPS KAKENHI Grant, Grant-in-Aid for Scientific Research (C) (22K08404 to Y.I.).

Acknowledgments

Y.I. would like to thank the following people, without whom he would not have been able to complete this proposal. Current or previous members in the Department of Dermatology at the University of Tsukuba, especially Tatsuya Ogawa for his diligence, Yasuhiro Fujisawa, Naoko Okiyama, Yoshiyuki Nakamura, and Noriko Kubota for helpful discussions and financial support. The Department of Dermatology at Osaka University members, especially Rei Watanabe and Manabu Fujimoto, for logistic support. The authors would like to display a feeling of fellowship between the Society for Investigative Dermatology and the Japanese Society for Investigative Dermatology.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Abbreviations

ACDallergic contact dermatitis
ADatopic dermatitis
ARCIautosomal recessive congenital ichthyosis
Argarginine
Ca2+calcium
CARD14caspase recruitment domain family member 14
CDSNcorneodesmosin
CEcornified cell envelope
CIEcongenital ichthyosiform erythroderma
CLEcorneocyte lipid envelopes
DAMPdamage-associated molecular pattern
DCdendritic cell
DETCdendritic epidermal T cell
DKOdouble knockout
DKSdesmosome-keratin scaffold
DLNdraining lymph node
DMBA7,12-dimethylbenz(a)anthracene
DNrf2dominant-negative Nrf2
DSG1desmoglein 1
ECADE-cadherin
EDCepidermal differentiation complex
EDUepidermal differentiation unit
EpiCAMepithelial cell adhesion molecule
EPUepidermal proliferation unit
FLGfilaggrin
FTTfailure to thrive
HIharlequin ichthyoses
IFEinterfollicular epidermis
ILinterleukin
ILCichthyosis linearis circumflexa
KCkeratinocyte
KEAP1Kelch-like erythroid cell-derived protein with the cap ‘n’ collar homology-associated protein 1
KGkeratohyalin granules
KLKkallikrein
LCLangerhans cell
LCElate cornified cell envelope proteins
LEKTILympho-epithelial Kazal-type related inhibitor type 5
LGlamellar granule
LIlamellar ichthyoses
LKOloricrin-knockout
LOFloss-of-function
LORloricrin
NMFnatural moisturizing factors
NRF2nuclear factor erythroid 2-related factor 2
NSNetherton syndrome
OMIMOnline Mendelian Inheritance in Man
PKKpalmoplantar keratoderma
PSSpeeling skin syndrome
redoxreduction and oxidation
SAMskin dermatitis, multiple severe allergies, and metabolic wasting
SBstratum basale
SCstratum corneum
SCCEstratum corneum chymotryptic enzyme
SCTEstratum corneum tryptic enzyme
SGstratum granulosum
SPINK5serine protease inhibitor of Kasal-type 5
SPRRsmall proline-rich proteins
TEWLtransepidermal water loss
TGF-βtransforming growth factor-beta
TGMtransglutaminase
THT helper
TIRtoll-interleukin receptor
TJtight junction
UVultraviolet
VAvitamin A
VSVohwinkel syndrome
WTwild type

References

  1. Autumn, K.; Liang, Y.A.; Hsieh, S.T.; Zesch, W.; Chan, W.P.; Kenny, T.W.; Fearing, R.; Full, R.J. Adhesive force of a single gecko foot-hair. Nature 2000, 405, 681–685. [Google Scholar] [CrossRef]
  2. Ito, T.; Ito, N.; Saathoff, M.; Stampachiacchiere, B.; Bettermann, A.; Bulfone-Paus, S.; Takigawa, M.; Nickoloff, B.J.; Paus, R. Immunology of the human nail apparatus: The nail matrix is a site of relative immune privilege. J. Investig. Dermatol. 2005, 125, 1139–1148. [Google Scholar] [CrossRef]
  3. Eckhart, L.; Lippens, S.; Tschachler, E.; Declercq, W. Cell death by cornification. Biochim. Biophys. Acta BBA Mol. Cell Res. 2013, 1833, 3471–3480. [Google Scholar] [CrossRef]
  4. Rice, R.H.; Durbin-Johnson, B.P.; Mann, S.M.; Salemi, M.; Urayama, S.; Rocke, D.M.; Phinney, B.S.; Sundberg, J.P. Corneocyte proteomics: Applications to skin biology and dermatology. Exp. Dermatol. 2018, 27, 931–938. [Google Scholar] [CrossRef] [Green Version]
  5. Rice, R.H.; Green, H. The cornified envelope of terminally differentiated human epidermal keratinocytes consists of cross-linked protein. Cell 1977, 11, 417–422. [Google Scholar] [CrossRef] [Green Version]
  6. Van Scott, E.; Flesch, P. Sulfhydryl and disulfide in keratinization. Science 1954, 119, 70–71. [Google Scholar] [CrossRef] [PubMed]
  7. Fukuyama, K.; Epstein, W.L. Sulfur-containing proteins and epidermal keratinization. J. Cell Biol. 1969, 40, 830–838. [Google Scholar] [CrossRef] [PubMed]
  8. Sharpen, J.D.A.; Dolan, B.; Nystrom, E.E.L.; Birchenough, G.M.H.; Arike, L.; Martinez-Abad, B.; Johansson, M.E.V.; Hansson, G.C.; Recktenwald, C.V. Transglutaminase 3 crosslinks the secreted gel-forming mucus component mucin-2 and stabilizes the colonic mucus layer. Nat. Commun. 2022, 13, 45. [Google Scholar] [CrossRef] [PubMed]
  9. Herrmann, A.; Davies, J.R.; Lindell, G.; Martensson, S.; Packer, N.H.; Swallow, D.M.; Carlstedt, I. Studies on the "insoluble" glycoprotein complex from human colon. Identification of reduction-insensitive muc2 oligomers and c-terminal cleavage. J. Biol. Chem. 1999, 274, 15828–15836. [Google Scholar] [CrossRef] [Green Version]
  10. Matoltsy, A.G.; Balsamo, C.A. A study of the components of the cornified epithelium of human skin. J. Biophys. Biochem. Cytol. 1955, 1, 339–360. [Google Scholar] [CrossRef] [Green Version]
  11. Teumer, J.; Green, H. Divergent evolution of part of the involucrin gene in the hominoids: Unique intragenic duplications in the gorilla and human. Proc. Natl. Acad. Sci. USA 1989, 86, 1283–1286. [Google Scholar] [CrossRef] [Green Version]
  12. Backendorf, C.; Hohl, D. A common origin for cornified envelope proteins? Nat. Genet. 1992, 2, 91. [Google Scholar] [CrossRef]
  13. Vandebergh, W.; Maex, M.; Bossuyt, F.; Van Bocxlaer, I. Recurrent functional divergence of early tetrapod keratins in amphibian toe pads and mammalian hair. Biol. Lett. 2013, 9, 20130051. [Google Scholar] [CrossRef] [Green Version]
  14. Alibardi, L. Adaptation to the land: The skin of reptiles in comparison to that of amphibians and endotherm amniotes. J. Exp. Zool. B Mol. Dev. Evol. 2003, 298, 12–41. [Google Scholar] [CrossRef]
  15. Federle, W.; Barnes, W.; Baumgartner, W.; Drechsler, P.; Smith, J. Wet but not slippery: Boundary friction in tree frog adhesive toe pads. J. R. Soc. Interface 2006, 3, 689–697. [Google Scholar] [CrossRef] [Green Version]
  16. Yokoyama, H.; Kudo, N.; Todate, M.; Shimada, Y.; Suzuki, M.; Tamura, K. Skin regeneration of amphibians: A novel model for skin regeneration as adults. Dev. Growth Differ. 2018, 60, 316–325. [Google Scholar] [CrossRef]
  17. Ehrlich, F.; Lachner, J.; Hermann, M.; Tschachler, E.; Eckhart, L. Convergent evolution of cysteine-rich keratins in hard skin appendages of terrestrial vertebrates. Mol. Biol. Evol. 2020, 37, 982–993. [Google Scholar] [CrossRef]
  18. Ehrlich, F.; Fischer, H.; Langbein, L.; Praetzel-Wunder, S.; Ebner, B.; Figlak, K.; Weissenbacher, A.; Sipos, W.; Tschachler, E.; Eckhart, L. Differential evolution of the epidermal keratin cytoskeleton in terrestrial and aquatic mammals. Mol. Biol. Evol. 2019, 36, 328–340. [Google Scholar] [CrossRef] [Green Version]
  19. Eckhart, L.; Ehrlich, F. Evolution of trichocyte keratins. Hair Fibre Proteins Struct. Dev. 2018, 1054, 33–45. [Google Scholar]
  20. Strasser, B.; Mlitz, V.; Fischer, H.; Tschachler, E.; Eckhart, L. Comparative genomics reveals conservation of filaggrin and loss of caspase-14 in dolphins. Exp. Dermatol. 2015, 24, 365–369. [Google Scholar] [CrossRef] [Green Version]
  21. Strasser, B.; Mlitz, V.; Hermann, M.; Rice, R.H.; Eigenheer, R.A.; Alibardi, L.; Tschachler, E.; Eckhart, L. Evolutionary origin and diversification of epidermal barrier proteins in amniotes. Mol. Biol. Evol. 2014, 31, 3194–3205. [Google Scholar] [CrossRef] [Green Version]
  22. Takeuchi, O.; Akira, S. Pattern recognition receptors and inflammation. Cell 2010, 140, 805–820. [Google Scholar] [CrossRef] [Green Version]
  23. Kemeny, L. The golden ages of inflammatory skin diseases: Skyrocketing developments in the therapy of psoriasis and atopic dermatitis. J. Eur. Acad. Dermatol. Venereol. 2021, 35, 2239–2240. [Google Scholar] [CrossRef]
  24. Kurimoto, K.; Yabuta, Y.; Ohinata, Y.; Saitou, M. Global single-cell cdna amplification to provide a template for representative high-density oligonucleotide microarray analysis. Nat. Protoc. 2007, 2, 739–752. [Google Scholar] [CrossRef]
  25. Ge, Y.; Gomez, N.C.; Adam, R.C.; Nikolova, M.; Yang, H.; Verma, A.; Lu, C.P.; Polak, L.; Yuan, S.; Elemento, O.; et al. Stem cell lineage infidelity drives wound repair and cancer. Cell 2017, 169, 636–650.e14. [Google Scholar] [CrossRef]
  26. Nakamizo, S.; Dutertre, C.A.; Khalilnezhad, A.; Zhang, X.M.; Lim, S.; Lum, J.; Koh, G.; Foong, C.; Yong, P.J.A.; Tan, K.J.; et al. Single-cell analysis of human skin identifies cd14+ type 3 dendritic cells co-producing il1b and il23a in psoriasis. J. Exp. Med. 2021, 218, e20202345. [Google Scholar] [CrossRef]
  27. Palmer, C.N.; Irvine, A.D.; Terron-Kwiatkowski, A.; Zhao, Y.; Liao, H.; Lee, S.P.; Goudie, D.R.; Sandilands, A.; Campbell, L.E.; Smith, F.J. Common loss-of-function variants of the epidermal barrier protein filaggrin are a major predisposing factor for atopic dermatitis. Nat. Genet. 2006, 38, 441. [Google Scholar] [CrossRef]
  28. Jordan, C.T.; Cao, L.; Roberson, E.D.; Duan, S.; Helms, C.A.; Nair, R.P.; Duffin, K.C.; Stuart, P.E.; Goldgar, D.; Hayashi, G.; et al. Rare and common variants in card14, encoding an epidermal regulator of nf-kappab, in psoriasis. Am. J. Hum. Genet. 2012, 90, 796–808. [Google Scholar] [CrossRef] [Green Version]
  29. Lieberman, D. The Story of the Human Body: Evolution, Health, and Disease; Vintage: New York, NY, USA, 2014. [Google Scholar]
  30. Ishitsuka, Y.; Ogawa, T.; Roop, D. The keap1/nrf2 signaling pathway in keratinization. Antioxidants 2020, 9, 751. [Google Scholar] [CrossRef]
  31. Ishitsuka, Y.; Roop, D.R. Loricrin: Past, present, and future. Int. J. Mol. Sci. 2020, 21, 2271. [Google Scholar] [CrossRef] [Green Version]
  32. Ishitsuka, Y.; Roop, D.R.; Ogawa, T. “Structural imprinting” of the cutaneous immune effector function. Tissue Barriers 2020, 9, 1851561. [Google Scholar] [CrossRef] [PubMed]
  33. Ishitsuka, Y.; Roop, D.R. The epidermis: Redox governor of health and diseases. Antioxidants 2022, 11, 47. [Google Scholar] [CrossRef] [PubMed]
  34. Rompolas, P.; Mesa, K.R.; Kawaguchi, K.; Park, S.; Gonzalez, D.; Brown, S.; Boucher, J.; Klein, A.M.; Greco, V. Spatiotemporal coordination of stem cell commitment during epidermal homeostasis. Science 2016, 352, 1471–1474. [Google Scholar] [CrossRef] [Green Version]
  35. Steven, A.C.; Bisher, M.E.; Roop, D.R.; Steinert, P.M. Biosynthetic pathways of filaggrin and loricrin--two major proteins expressed by terminally differentiated epidermal keratinocytes. J. Struct. Biol. 1990, 104, 150–162. [Google Scholar] [CrossRef]
  36. Matsui, T.; Amagai, M. Dissecting the formation, structure and barrier function of the stratum corneum. Int. Immunol. 2015, 27, 269–280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Matsui, T.; Kadono-Maekubo, N.; Suzuki, Y.; Furuichi, Y.; Shiraga, K.; Sasaki, H.; Ishida, A.; Takahashi, S.; Okada, T.; Toyooka, K.; et al. A unique mode of keratinocyte death requires intracellular acidification. Proc. Natl. Acad. Sci. USA 2021, 118, e2020722118. [Google Scholar] [CrossRef]
  38. Murata, T.; Honda, T.; Egawa, G.; Yamamoto, Y.; Ichijo, R.; Toyoshima, F.; Dainichi, T.; Kabashima, K. Transient elevation of cytoplasmic calcium ion concentration at a single cell level precedes morphological changes of epidermal keratinocytes during cornification. Sci. Rep. 2018, 8, 6610. [Google Scholar] [CrossRef]
  39. Akinduro, O.; Sully, K.; Patel, A.; Robinson, D.J.; Chikh, A.; McPhail, G.; Braun, K.M.; Philpott, M.P.; Harwood, C.A.; Byrne, C.; et al. Constitutive autophagy and nucleophagy during epidermal differentiation. J. Investig. Dermatol. 2016, 136, 1460–1470. [Google Scholar] [CrossRef] [Green Version]
  40. Furio, L.; de Veer, S.; Jaillet, M.; Briot, A.; Robin, A.; Deraison, C.; Hovnanian, A. Transgenic kallikrein 5 mice reproduce major cutaneous and systemic hallmarks of netherton syndrome. J. Exp. Med. 2014, 211, 499–513. [Google Scholar] [CrossRef] [Green Version]
  41. Matsui, T.; Miyamoto, K.; Kubo, A.; Kawasaki, H.; Ebihara, T.; Hata, K.; Tanahashi, S.; Ichinose, S.; Imoto, I.; Inazawa, J.; et al. Saspase regulates stratum corneum hydration through profilaggrin-to-filaggrin processing. EMBO Mol. Med. 2011, 3, 320–333. [Google Scholar] [CrossRef]
  42. Denecker, G.; Hoste, E.; Gilbert, B.; Hochepied, T.; Ovaere, P.; Lippens, S.; Van den Broecke, C.; Van Damme, P.; D’Herde, K.; Hachem, J.P.; et al. Caspase-14 protects against epidermal uvb photodamage and water loss. Nat. Cell Biol. 2007, 9, 666–674. [Google Scholar] [CrossRef]
  43. Descargues, P.; Deraison, C.; Bonnart, C.; Kreft, M.; Kishibe, M.; Ishida-Yamamoto, A.; Elias, P.; Barrandon, Y.; Zambruno, G.; Sonnenberg, A.; et al. Spink5-deficient mice mimic netherton syndrome through degradation of desmoglein 1 by epidermal protease hyperactivity. Nat. Genet. 2005, 37, 56–65. [Google Scholar] [CrossRef]
  44. Schwartz, D.R.; Homanics, G.E.; Hoyt, D.G.; Klein, E.; Abernethy, J.; Lazo, J.S. The neutral cysteine protease bleomycin hydrolase is essential for epidermal integrity and bleomycin resistance. Proc. Natl. Acad. Sci. USA 1999, 96, 4680–4685. [Google Scholar] [CrossRef] [Green Version]
  45. Imakado, S.; Bickenbach, J.R.; Bundman, D.S.; Rothnagel, J.A.; Attar, P.S.; Wang, X.J.; Walczak, V.R.; Wisniewski, S.; Pote, J.; Gordon, J.S.; et al. Targeting expression of a dominant-negative retinoic acid receptor mutant in the epidermis of transgenic mice results in loss of barrier function. Genes. Dev. 1995, 9, 317–329. [Google Scholar] [CrossRef] [Green Version]
  46. Xie, Z.; Komuves, L.; Yu, Q.C.; Elalieh, H.; Ng, D.C.; Leary, C.; Chang, S.; Crumrine, D.; Yoshizawa, T.; Kato, S.; et al. Lack of the vitamin d receptor is associated with reduced epidermal differentiation and hair follicle growth. J. Investig. Dermatol. 2002, 118, 11–16. [Google Scholar] [CrossRef] [Green Version]
  47. Ishida-Yamamoto, A.; Simon, M.; Kishibe, M.; Miyauchi, Y.; Takahashi, H.; Yoshida, S.; O’Brien, T.J.; Serre, G.; Iizuka, H. Epidermal lamellar granules transport different cargoes as distinct aggregates. J. Investig. Dermatol. 2004, 122, 1137–1144. [Google Scholar] [CrossRef] [Green Version]
  48. Raymond, A.A.; Gonzalez de Peredo, A.; Stella, A.; Ishida-Yamamoto, A.; Bouyssie, D.; Serre, G.; Monsarrat, B.; Simon, M. Lamellar bodies of human epidermis: Proteomics characterization by high throughput mass spectrometry and possible involvement of clip-170 in their trafficking/secretion. Mol. Cell Proteom. 2008, 7, 2151–2175. [Google Scholar] [CrossRef] [Green Version]
  49. Zuo, Y.; Zhuang, D.Z.; Han, R.; Isaac, G.; Tobin, J.J.; McKee, M.; Welti, R.; Brissette, J.L.; Fitzgerald, M.L.; Freeman, M.W. Abca12 maintains the epidermal lipid permeability barrier by facilitating formation of ceramide linoleic esters. J. Biol. Chem. 2008, 283, 36624–36635. [Google Scholar] [CrossRef] [Green Version]
  50. Kalinin, A.; Marekov, L.N.; Steinert, P.M. Assembly of the epidermal cornified cell envelope. J. Cell Sci. 2001, 114, 3069–3070. [Google Scholar] [CrossRef]
  51. Crumrine, D.; Khnykin, D.; Krieg, P.; Man, M.Q.; Celli, A.; Mauro, T.M.; Wakefield, J.S.; Menon, G.; Mauldin, E.; Miner, J.H.; et al. Mutations in recessive congenital ichthyoses illuminate the origin and functions of the corneocyte lipid envelope. J. Investig. Dermatol. 2019, 139, 760–768. [Google Scholar] [CrossRef] [Green Version]
  52. Harding, C.R.; Aho, S.; Bosko, C.A. Filaggrin—Revisited. Int. J. Cosmet. Sci. 2013, 35, 412–423. [Google Scholar] [CrossRef] [PubMed]
  53. Bognar, P.; Nemeth, I.; Mayer, B.; Haluszka, D.; Wikonkal, N.; Ostorhazi, E.; John, S.; Paulsson, M.; Smyth, N.; Pasztoi, M.; et al. Reduced inflammatory threshold indicates skin barrier defect in transglutaminase 3 knockout mice. J. Investig. Dermatol. 2014, 134, 105–111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Matsuki, M.; Yamashita, F.; Ishida-Yamamoto, A.; Yamada, K.; Kinoshita, C.; Fushiki, S.; Ueda, E.; Morishima, Y.; Tabata, K.; Yasuno, H.; et al. Defective stratum corneum and early neonatal death in mice lacking the gene for transglutaminase 1 (keratinocyte transglutaminase). Proc. Natl. Acad. Sci. USA 1998, 95, 1044–1049. [Google Scholar] [CrossRef] [Green Version]
  55. Ishitsuka, Y.; Ogawa, T.; Nakamura, Y.; Kubota, N.; Fujisawa, Y.; Watanabe, R.; Okiyama, N.; Fujimoto, M.; Roop, D.R.; Ishida-Yamamoto, A. Loricrin and nrf2 coordinate cornification. JID Innov. 2021, 2, 100065. [Google Scholar] [CrossRef] [PubMed]
  56. Ishitsuka, Y.; Roop, D.R. Loricrin confers photoprotective function against uvb in corneocytes. J. Investig. Dermatol. 2018, 138, 2684–2687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Rice, R.H.; Durbin-Johnson, B.P.; Ishitsuka, Y.; Salemi, M.; Phinney, B.S.; Rocke, D.M.; Roop, D.R. Proteomic analysis of loricrin knockout mouse epidermis. J. Proteome Res. 2016, 15, 2560–2566. [Google Scholar] [CrossRef] [PubMed]
  58. Huebner, A.J.; Dai, D.; Morasso, M.; Schmidt, E.E.; Schafer, M.; Werner, S.; Roop, D.R. Amniotic fluid activates the nrf2/keap1 pathway to repair an epidermal barrier defect in utero. Dev. Cell 2012, 23, 1238–1246. [Google Scholar] [CrossRef] [Green Version]
  59. Koch, P.J.; de Viragh, P.A.; Scharer, E.; Bundman, D.; Longley, M.A.; Bickenbach, J.; Kawachi, Y.; Suga, Y.; Zhou, Z.; Huber, M.; et al. Lessons from loricrin-deficient mice: Compensatory mechanisms maintaining skin barrier function in the absence of a major cornified envelope protein. J. Cell Biol. 2000, 151, 389–400. [Google Scholar] [CrossRef]
  60. Steinert, P.M.; Marekov, L.N. The proteins elafin, filaggrin, keratin intermediate filaments, loricrin, and small proline-rich proteins 1 and 2 are isodipeptide cross-linked components of the human epidermal cornified cell envelope. J. Biol. Chem. 1995, 270, 17702–17711. [Google Scholar] [CrossRef] [Green Version]
  61. Hohl, D.; Mehrel, T.; Lichti, U.; Turner, M.L.; Roop, D.R.; Steinert, P.M. Characterization of human loricrin. Structure and function of a new class of epidermal cell envelope proteins. J. Biol. Chem. 1991, 266, 6626–6636. [Google Scholar] [CrossRef]
  62. Mehrel, T.; Hohl, D.; Rothnagel, J.A.; Longley, M.A.; Bundman, D.; Cheng, C.; Lichti, U.; Bisher, M.E.; Steven, A.C.; Steinert, P.M.; et al. Identification of a major keratinocyte cell envelope protein, loricrin. Cell 1990, 61, 1103–1112. [Google Scholar] [CrossRef]
  63. Nemes, Z.; Marekov, L.N.; Fesus, L.; Steinert, P.M. A novel function for transglutaminase 1: Attachment of long-chain omega-hydroxyceramides to involucrin by ester bond formation. Proc. Natl. Acad. Sci. USA 1999, 96, 8402–8407. [Google Scholar] [CrossRef] [Green Version]
  64. Ishida-Yamamoto, A.; Eady, R.A.; Watt, F.M.; Roop, D.R.; Hohl, D.; Iizuka, H. Immunoelectron microscopic analysis of cornified cell envelope formation in normal and psoriatic epidermis. J. Histochem. Cytochem. 1996, 44, 167–175. [Google Scholar] [CrossRef] [Green Version]
  65. Ishida-Yamamoto, A.; Tanaka, H.; Nakane, H.; Takahashi, H.; Iizuka, H. Antigen retrieval of loricrin epitopes at desmosomal areas of cornified cell envelopes: An immunoelectron microscopic analysis. Exp. Dermatol. 1999, 8, 402–406. [Google Scholar] [CrossRef]
  66. Hatzfeld, M.; Keil, R.; Magin, T.M. Desmosomes and intermediate filaments: Their consequences for tissue mechanics. Cold Spring Harb. Perspect. Biol. 2017, 9, a029157. [Google Scholar] [CrossRef] [Green Version]
  67. Steinert, P.M.; Cantieri, J.S.; Teller, D.C.; Lonsdale-Eccles, J.D.; Dale, B.A. Characterization of a class of cationic proteins that specifically interact with intermediate filaments. Proc. Natl. Acad. Sci. USA 1981, 78, 4097–4101. [Google Scholar] [CrossRef] [Green Version]
  68. Moniaga, C.S.; Egawa, G.; Kawasaki, H.; Hara-Chikuma, M.; Honda, T.; Tanizaki, H.; Nakajima, S.; Otsuka, A.; Matsuoka, H.; Kubo, A.; et al. Flaky tail mouse denotes human atopic dermatitis in the steady state and by topical application with dermatophagoides pteronyssinus extract. Am. J. Pathol. 2010, 176, 2385–2393. [Google Scholar] [CrossRef]
  69. Kawasaki, H.; Nagao, K.; Kubo, A.; Hata, T.; Shimizu, A.; Mizuno, H.; Yamada, T.; Amagai, M. Altered stratum corneum barrier and enhanced percutaneous immune responses in filaggrin-null mice. J. Allergy Clin. Immunol. 2012, 129, 1538–1546.e6. [Google Scholar] [CrossRef] [Green Version]
  70. Novak, N.; Baurecht, H.; Schafer, T.; Rodriguez, E.; Wagenpfeil, S.; Klopp, N.; Heinrich, J.; Behrendt, H.; Ring, J.; Wichmann, E.; et al. Loss-of-function mutations in the filaggrin gene and allergic contact sensitization to nickel. J. Investig. Dermatol. 2008, 128, 1430–1435. [Google Scholar] [CrossRef] [Green Version]
  71. Kobayashi, T.; Glatz, M.; Horiuchi, K.; Kawasaki, H.; Akiyama, H.; Kaplan, D.H.; Kong, H.H.; Amagai, M.; Nagao, K. Dysbiosis and staphylococcus aureus colonization drives inflammation in atopic dermatitis. Immunity 2015, 42, 756–766. [Google Scholar] [CrossRef] [Green Version]
  72. Ishida-Yamamoto, A.; Igawa, S.; Kishibe, M. Order and disorder in corneocyte adhesion. J. Dermatol. 2011, 38, 645–654. [Google Scholar] [CrossRef] [Green Version]
  73. Ishida-Yamamoto, A.; Igawa, S.; Kishibe, M. Molecular basis of the skin barrier structures revealed by electron microscopy. Exp. Dermatol. 2018, 27, 841–846. [Google Scholar] [CrossRef] [Green Version]
  74. Brattsand, M.; Egelrud, T. Purification, molecular cloning, and expression of a human stratum corneum trypsin-like serine protease with possible function in desquamation. J. Biol. Chem. 1999, 274, 30033–30040. [Google Scholar] [CrossRef] [Green Version]
  75. Godsel, L.M.; Roth-Carter, Q.R.; Koetsier, J.L.; Tsoi, L.C.; Huffine, A.L.; Broussard, J.A.; Fitz, G.N.; Lloyd, S.M.; Kweon, J.; Burks, H.E.; et al. Translational implications of th17-skewed inflammation due to genetic deficiency of a cadherin stress sensor. J. Clin. Investig. 2021, 132, e144363. [Google Scholar] [CrossRef]
  76. Kugelmann, D.; Radeva, M.Y.; Spindler, V.; Waschke, J. Desmoglein 1 deficiency causes lethal skin blistering. J. Investig. Dermatol. 2019, 139, 1596–1599.e2. [Google Scholar] [CrossRef]
  77. Leclerc, E.A.; Huchenq, A.; Mattiuzzo, N.R.; Metzger, D.; Chambon, P.; Ghyselinck, N.B.; Serre, G.; Jonca, N.; Guerrin, M. Corneodesmosin gene ablation induces lethal skin-barrier disruption and hair-follicle degeneration related to desmosome dysfunction. J. Cell Sci. 2009, 122, 2699–2709. [Google Scholar] [CrossRef] [Green Version]
  78. Matsumoto, M.; Zhou, Y.; Matsuo, S.; Nakanishi, H.; Hirose, K.; Oura, H.; Arase, S.; Ishida-Yamamoto, A.; Bando, Y.; Izumi, K.; et al. Targeted deletion of the murine corneodesmosin gene delineates its essential role in skin and hair physiology. Proc. Natl. Acad. Sci. USA 2008, 105, 6720–6724. [Google Scholar] [CrossRef] [Green Version]
  79. Chavanas, S.; Bodemer, C.; Rochat, A.; Hamel-Teillac, D.; Ali, M.; Irvine, A.D.; Bonafe, J.L.; Wilkinson, J.; Taieb, A.; Barrandon, Y.; et al. Mutations in spink5, encoding a serine protease inhibitor, cause netherton syndrome. Nat. Genet. 2000, 25, 141–142. [Google Scholar] [CrossRef]
  80. Samuelov, L.; Sarig, O.; Harmon, R.M.; Rapaport, D.; Ishida-Yamamoto, A.; Isakov, O.; Koetsier, J.L.; Gat, A.; Goldberg, I.; Bergman, R.; et al. Desmoglein 1 deficiency results in severe dermatitis, multiple allergies and metabolic wasting. Nat. Genet. 2013, 45, 1244–1248. [Google Scholar] [CrossRef] [Green Version]
  81. Samuelov, L.; Sprecher, E. Peeling off the genetics of atopic dermatitis-like congenital disorders. J. Allergy Clin. Immunol. 2014, 134, 808–815. [Google Scholar] [CrossRef]
  82. Hadj-Rabia, S.; Baala, L.; Vabres, P.; Hamel-Teillac, D.; Jacquemin, E.; Fabre, M.; Lyonnet, S.; De Prost, Y.; Munnich, A.; Hadchouel, M.; et al. Claudin-1 gene mutations in neonatal sclerosing cholangitis associated with ichthyosis: A tight junction disease. Gastroenterology 2004, 127, 1386–1390. [Google Scholar] [CrossRef] [PubMed]
  83. Hamosh, A.; Scott, A.F.; Amberger, J.S.; Bocchini, C.A.; McKusick, V.A. Online mendelian inheritance in man (omim), a knowledgebase of human genes and genetic disorders. Nucleic Acids Res. 2005, 33, D514–D517. [Google Scholar] [CrossRef] [PubMed]
  84. Maestrini, E.; Monaco, A.P.; McGrath, J.A.; Ishida-Yamamoto, A.; Camisa, C.; Hovnanian, A.; Weeks, D.E.; Lathrop, M.; Uitto, J.; Christiano, A.M. A molecular defect in loricrin, the major component of the cornified cell envelope, underlies vohwinkel’s syndrome. Nat. Genet. 1996, 13, 70–77. [Google Scholar] [CrossRef] [PubMed]
  85. Takeichi, T.; Akiyama, M. Inherited ichthyosis: Non-syndromic forms. J. Dermatol. 2016, 43, 242–251. [Google Scholar] [CrossRef]
  86. Takeichi, T.; Hirabayashi, T.; Miyasaka, Y.; Kawamoto, A.; Okuno, Y.; Taguchi, S.; Tanahashi, K.; Murase, C.; Takama, H.; Tanaka, K.; et al. Sdr9c7 catalyzes critical dehydrogenation of acylceramides for skin barrier formation. J. Clin. Investig. 2020, 130, 890–903. [Google Scholar] [CrossRef]
  87. Heinz, L.; Kim, G.J.; Marrakchi, S.; Christiansen, J.; Turki, H.; Rauschendorf, M.A.; Lathrop, M.; Hausser, I.; Zimmer, A.D.; Fischer, J. Mutations in sult2b1 cause autosomal-recessive congenital ichthyosis in humans. Am. J. Hum. Genet. 2017, 100, 926–939. [Google Scholar] [CrossRef] [Green Version]
  88. Ohno, Y.; Nakamichi, S.; Ohkuni, A.; Kamiyama, N.; Naoe, A.; Tsujimura, H.; Yokose, U.; Sugiura, K.; Ishikawa, J.; Akiyama, M.; et al. Essential role of the cytochrome p450 cyp4f22 in the production of acylceramide, the key lipid for skin permeability barrier formation. Proc. Natl. Acad. Sci. USA 2015, 112, 7707–7712. [Google Scholar] [CrossRef] [Green Version]
  89. Radner, F.P.; Marrakchi, S.; Kirchmeier, P.; Kim, G.J.; Ribierre, F.; Kamoun, B.; Abid, L.; Leipoldt, M.; Turki, H.; Schempp, W.; et al. Mutations in cers3 cause autosomal recessive congenital ichthyosis in humans. PLoS Genet. 2013, 9, e1003536. [Google Scholar] [CrossRef]
  90. Grall, A.; Guaguere, E.; Planchais, S.; Grond, S.; Bourrat, E.; Hausser, I.; Hitte, C.; Le Gallo, M.; Derbois, C.; Kim, G.J.; et al. Pnpla1 mutations cause autosomal recessive congenital ichthyosis in golden retriever dogs and humans. Nat. Genet. 2012, 44, 140–147. [Google Scholar] [CrossRef]
  91. Israeli, S.; Khamaysi, Z.; Fuchs-Telem, D.; Nousbeck, J.; Bergman, R.; Sarig, O.; Sprecher, E. A mutation in lipn, encoding epidermal lipase n, causes a late-onset form of autosomal-recessive congenital ichthyosis. Am. J. Hum. Genet. 2011, 88, 482–487. [Google Scholar] [CrossRef] [Green Version]
  92. Wajid, M.; Kurban, M.; Shimomura, Y.; Christiano, A.M. Nipal4/ichthyin is expressed in the granular layer of human epidermis and mutated in two pakistani families with autosomal recessive ichthyosis. Dermatology 2010, 220, 8–14. [Google Scholar] [CrossRef] [Green Version]
  93. Jobard, F.; Lefevre, C.; Karaduman, A.; Blanchet-Bardon, C.; Emre, S.; Weissenbach, J.; Ozguc, M.; Lathrop, M.; Prud’homme, J.F.; Fischer, J. Lipoxygenase-3 (aloxe3) and 12(r)-lipoxygenase (alox12b) are mutated in non-bullous congenital ichthyosiform erythroderma (ncie) linked to chromosome 17p13.1. Hum. Mol. Genet. 2002, 11, 107–113. [Google Scholar] [CrossRef] [Green Version]
  94. Akiyama, M.; Sugiyama-Nakagiri, Y.; Sakai, K.; McMillan, J.R.; Goto, M.; Arita, K.; Tsuji-Abe, Y.; Tabata, N.; Matsuoka, K.; Sasaki, R.; et al. Mutations in lipid transporter abca12 in harlequin ichthyosis and functional recovery by corrective gene transfer. J. Clin. Investig. 2005, 115, 1777–1784. [Google Scholar] [CrossRef] [Green Version]
  95. Huber, M.; Rettler, I.; Bernasconi, K.; Frenk, E.; Lavrijsen, S.P.; Ponec, M.; Bon, A.; Lautenschlager, S.; Schorderet, D.F.; Hohl, D. Mutations of keratinocyte transglutaminase in lamellar ichthyosis. Science 1995, 267, 525–528. [Google Scholar] [CrossRef]
  96. Kirchmeier, P.; Zimmer, A.; Bouadjar, B.; Rosler, B.; Fischer, J. Whole-exome-sequencing reveals small deletions in casp14 in patients with autosomal recessive inherited ichthyosis. Acta Derm. Venereol. 2017, 97, 102–104. [Google Scholar] [CrossRef] [Green Version]
  97. Basel-Vanagaite, L.; Attia, R.; Ishida-Yamamoto, A.; Rainshtein, L.; Amitai, D.B.; Lurie, R.; Pasmanik-Chor, M.; Indelman, M.; Zvulunov, A.; Saban, S. Autosomal recessive ichthyosis with hypotrichosis caused by a mutation in st14, encoding type ii transmembrane serine protease matriptase. Am. J. Hum. Genet. 2007, 80, 467–477. [Google Scholar] [CrossRef] [Green Version]
  98. Oji, V.; Eckl, K.M.; Aufenvenne, K.; Natebus, M.; Tarinski, T.; Ackermann, K.; Seller, N.; Metze, D.; Nurnberg, G.; Folster-Holst, R.; et al. Loss of corneodesmosin leads to severe skin barrier defect, pruritus, and atopy: Unraveling the peeling skin disease. Am. J. Hum. Genet. 2010, 87, 274–281. [Google Scholar] [CrossRef] [Green Version]
  99. Comel, M. Ichthyosis linearis circumflexa. Dermatology 1949, 98, 133–136. [Google Scholar] [CrossRef]
  100. Wallace, L.; Roberts-Thompson, L.; Reichelt, J. Deletion of k1/k10 does not impair epidermal stratification but affects desmosomal structure and nuclear integrity. J. Cell Sci. 2012, 125, 1750–1758. [Google Scholar] [CrossRef] [Green Version]
  101. Boyle, C.J.; Plotczyk, M.; Villalta, S.F.; Patel, S.; Hettiaratchy, S.; Masouros, S.D.; Masen, M.A.; Higgins, C.A. Morphology and composition play distinct and complementary roles in the tolerance of plantar skin to mechanical load. Sci. Adv. 2019, 5, eaay0244. [Google Scholar] [CrossRef] [Green Version]
  102. Ishida-Yamamoto, A.; McGrath, J.A.; Lam, H.; Iizuka, H.; Friedman, R.A.; Christiano, A.M. The molecular pathology of progressive symmetric erythrokeratoderma: A frameshift mutation in the loricrin gene and perturbations in the cornified cell envelope. Am. J. Hum. Genet. 1997, 61, 581–589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Suga, Y.; Jarnik, M.; Attar, P.S.; Longley, M.A.; Bundman, D.; Steven, A.C.; Koch, P.J.; Roop, D.R. Transgenic mice expressing a mutant form of loricrin reveal the molecular basis of the skin diseases, vohwinkel syndrome and progressive symmetric erythrokeratoderma. J. Cell Biol. 2000, 151, 401–412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Paller, A.S.; Renert-Yuval, Y.; Suprun, M.; Esaki, H.; Oliva, M.; Huynh, T.N.; Ungar, B.; Kunjravia, N.; Friedland, R.; Peng, X.; et al. An il-17-dominant immune profile is shared across the major orphan forms of ichthyosis. J. Allergy Clin. Immunol. 2017, 139, 152–165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Altman, J.; Stroud, J. Neterton’s syndrome and ichthyosis linearis circumflexa. Arch. Dermatol. 1969, 100, 550–558. [Google Scholar] [CrossRef]
  106. Paller, A.S. Profiling immune expression to consider repurposing therapeutics for the ichthyoses. J. Investig. Dermatol. 2019, 139, 535–540. [Google Scholar] [CrossRef] [Green Version]
  107. McKenzie, B.S.; Kastelein, R.A.; Cua, D.J. Understanding the il-23-il-17 immune pathway. Trends Immunol. 2006, 27, 17–23. [Google Scholar] [CrossRef]
  108. Pinkus, H. Psoriasiform tissue reactions. Aust. J. Dermatol. 1965, 8, 31–35. [Google Scholar] [CrossRef]
  109. Ouchi, T.; Kubo, A.; Yokouchi, M.; Adachi, T.; Kobayashi, T.; Kitashima, D.Y.; Fujii, H.; Clausen, B.E.; Koyasu, S.; Amagai, M.; et al. Langerhans cell antigen capture through tight junctions confers preemptive immunity in experimental staphylococcal scalded skin syndrome. J. Exp. Med. 2011, 208, 2607–2613. [Google Scholar] [CrossRef] [Green Version]
  110. Swamy, M.; Jamora, C.; Havran, W.; Hayday, A. Epithelial decision makers: In search of the ‘epimmunome’. Nat. Immunol. 2010, 11, 656–665. [Google Scholar] [CrossRef] [Green Version]
  111. Matzinger, P.; Kamala, T. Tissue-based class control: The other side of tolerance. Nat. Rev. Immunol. 2011, 11, 221–230. [Google Scholar] [CrossRef]
  112. Dainichi, T.; Kitoh, A.; Otsuka, A.; Nakajima, S.; Nomura, T.; Kaplan, D.H.; Kabashima, K. The epithelial immune microenvironment (eime) in atopic dermatitis and psoriasis. Nat. Immunol. 2018, 19, 1286–1298. [Google Scholar] [CrossRef]
  113. Barbieux, C.; Des Claustres, M.B.; De la Brassinne, M.; Bricteux, G.; Bagot, M.; Bourrat, E.; Hovnanian, A. Duality of netherton syndrome manifestations and response to ixekizumab. J. Am. Acad. Dermatol. 2021, 84, 1476–1480. [Google Scholar] [CrossRef]
  114. Blanchard, S.K.; Prose, N.S. Successful use of secukinumab in Netherton syndrome. JAAD Case Rep. 2020, 6, 577–578. [Google Scholar] [CrossRef]
  115. Lefferdink, R.; Rangel, S.M.; Chima, M.; Ibler, E.; Pavel, A.B.; Kim, H.; Wu, B.; Abu-Zayed, H.; Wu, J.; Jackson, K.; et al. Secukinumab responses vary across the spectrum of congenital ichthyosis in adults. Arch. Dermatol. Res. 2022, 1–11. [Google Scholar] [CrossRef]
  116. Luchsinger, I.; Knöpfel, N.; Theiler, M.; des Claustres, M.B.; Barbieux, C.; Schwieger-Briel, A.; Brunner, C.; Donghi, D.; Buettcher, M.; Meier-Schiesser, B. Secukinumab therapy for Netherton syndrome. JAMA Dermatol. 2020, 156, 907–911. [Google Scholar] [CrossRef]
  117. Mischke, D.; Korge, B.P.; Marenholz, I.; Volz, A.; Ziegler, A. Genes encoding structural proteins of epidermal cornification and S100 calcium-binding proteins form a gene complex (“epidermal differentiation complex”) on human chromosome 1q21. J. Investig. Dermatol. 1996, 106, 989–992. [Google Scholar] [CrossRef] [Green Version]
  118. Ishitsuka, Y.; Huebner, A.J.; Rice, R.H.; Koch, P.J.; Speransky, V.V.; Steven, A.C.; Roop, D.R. Lce1 Family Members Are Nrf2-Target Genes that Are Induced to Compensate for the Loss of Loricrin. J. Investig. Dermatol. 2016, 136, 1656–1663. [Google Scholar] [CrossRef] [Green Version]
  119. Hu, Z.; Zhang, C.; Sifuentes-Dominguez, L.; Zarek, C.M.; Propheter, D.C.; Kuang, Z.; Wang, Y.; Pendse, M.; Ruhn, K.A.; Hassell, B.; et al. Small proline-rich protein 2A is a gut bactericidal protein deployed during helminth infection. Science 2021, 374, eabe6723. [Google Scholar] [CrossRef]
  120. Huynh, K.M.; Wong, A.C.; Wu, B.; Horschman, M.; Zhao, H.; Brooks, J.D. Sprr2f protects against renal injury by decreasing the level of reactive oxygen species in female mice. Am. J. Physiol. Ren. Physiol. 2020, 319, F876–F884. [Google Scholar] [CrossRef]
  121. Vermeij, W.P.; Alia, A.; Backendorf, C. ROS quenching potential of the epidermal cornified cell envelope. J. Investig. Dermatol. 2011, 131, 1435–1441. [Google Scholar] [CrossRef] [Green Version]
  122. Vermeij, W.P.; Backendorf, C. Skin cornification proteins provide global link between ros detoxification and cell migration during wound healing. PLoS ONE 2010, 5, e11957. [Google Scholar] [CrossRef] [Green Version]
  123. Gibbs, S.; Fijneman, R.; Wiegant, J.; van Kessel, A.G.; van De Putte, P.; Backendorf, C. Molecular characterization and evolution of the sprr family of keratinocyte differentiation markers encoding small proline-rich proteins. Genomics 1993, 16, 630–637. [Google Scholar] [CrossRef]
  124. Jackson, B.; Tilli, C.M.; Hardman, M.J.; Avilion, A.A.; MacLeod, M.C.; Ashcroft, G.S.; Byrne, C. Late cornified envelope family in differentiating epithelia—Response to calcium and ultraviolet irradiation. J. Investig. Dermatol. 2005, 124, 1062–1070. [Google Scholar] [CrossRef]
  125. De Cid, R.; Riveira-Munoz, E.; Zeeuwen, P.L.; Robarge, J.; Liao, W.; Dannhauser, E.N.; Giardina, E.; Stuart, P.E.; Nair, R.; Helms, C.; et al. Deletion of the late cornified envelope lce3b and lce3c genes as a susceptibility factor for psoriasis. Nat. Genet. 2009, 41, 211–215. [Google Scholar] [CrossRef] [Green Version]
  126. Niehues, H.; Tsoi, L.C.; van der Krieken, D.A.; Jansen, P.A.M.; Oortveld, M.A.W.; Rodijk-Olthuis, D.; van Vlijmen, I.; Hendriks, W.; Helder, R.W.; Bouwstra, J.A.; et al. Psoriasis-associated late cornified envelope (lce) proteins have antibacterial activity. J. Investig. Dermatol. 2017, 137, 2380–2388. [Google Scholar] [CrossRef] [Green Version]
  127. Baird, L.; Yamamoto, M. The molecular mechanisms regulating the keap1-nrf2 pathway. Mol. Cell. Biol. 2020, 40, e00099-20. [Google Scholar] [CrossRef]
  128. Kang, M.I.; Kobayashi, A.; Wakabayashi, N.; Kim, S.G.; Yamamoto, M. Scaffolding of keap1 to the actin cytoskeleton controls the function of nrf2 as key regulator of cytoprotective phase 2 genes. Proc. Natl. Acad. Sci. USA 2004, 101, 2046–2051. [Google Scholar] [CrossRef] [Green Version]
  129. Wakabayashi, N.; Itoh, K.; Wakabayashi, J.; Motohashi, H.; Noda, S.; Takahashi, S.; Imakado, S.; Kotsuji, T.; Otsuka, F.; Roop, D.R.; et al. Keap1-null mutation leads to postnatal lethality due to constitutive nrf2 activation. Nat. Genet. 2003, 35, 238–245. [Google Scholar] [CrossRef]
  130. Ogawa, T.; Ishitsuka, Y.; Fujimoto, M.; Nomura, T. Keap1 and epidermal differentiation: Psoriatic epidermis as a model. J. Cutan. Immunol. Allergy 2021, 4, 132–134. [Google Scholar] [CrossRef]
  131. Schafer, M.; Dutsch, S.; auf dem Keller, U.; Navid, F.; Schwarz, A.; Johnson, D.A.; Johnson, J.A.; Werner, S. Nrf2 establishes a glutathione-mediated gradient of uvb cytoprotection in the epidermis. Genes Dev. 2010, 24, 1045–1058. [Google Scholar] [CrossRef] [Green Version]
  132. Ogawa, T.; Ishitsuka, Y.; Nakamura, Y.; Kubota, N.; Saito, A.; Fujisawa, Y.; Watanabe, R.; Okiyama, N.; Suga, Y.; Roop, D.R.; et al. Nrf2 augments epidermal antioxidant defenses and promotes atopy. J. Immunol. 2020, 205, 907–914. [Google Scholar] [CrossRef] [PubMed]
  133. Xu, C.; Huang, M.T.; Shen, G.; Yuan, X.; Lin, W.; Khor, T.O.; Conney, A.H.; Kong, A.N. Inhibition of 7,12-dimethylbenz(a)anthracene-induced skin tumorigenesis in c57bl/6 mice by sulforaphane is mediated by nuclear factor e2-related factor 2. Cancer Res. 2006, 66, 8293–8296. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Li, J.; Ichikawa, T.; Villacorta, L.; Janicki, J.S.; Brower, G.L.; Yamamoto, M.; Cui, T. Nrf2 protects against maladaptive cardiac responses to hemodynamic stress. Arter. Thromb. Vasc. Biol. 2009, 29, 1843–1850. [Google Scholar] [CrossRef] [PubMed]
  135. Kawachi, Y.; Xu, X.; Taguchi, S.; Sakurai, H.; Nakamura, Y.; Ishii, Y.; Fujisawa, Y.; Furuta, J.; Takahashi, T.; Itoh, K.; et al. Attenuation of uvb-induced sunburn reaction and oxidative DNA damage with no alterations in uvb-induced skin carcinogenesis in nrf2 gene-deficient mice. J. Investig. Dermatol. 2008, 128, 1773–1779. [Google Scholar] [CrossRef] [Green Version]
  136. Schafer, M.; Farwanah, H.; Willrodt, A.H.; Huebner, A.J.; Sandhoff, K.; Roop, D.; Hohl, D.; Bloch, W.; Werner, S. Nrf2 links epidermal barrier function with antioxidant defense. EMBO Mol. Med. 2012, 4, 364–379. [Google Scholar] [CrossRef]
  137. Ogawa, T.; Ishitsuka, Y.; Nakamura, Y.; Watanabe, R.; Okiyama, N.; Fujisawa, Y.; Fujimoto, M.; Roop, D.R.; Nomura, T. Loricrin protects against chemical carcinogenesis. J. Investig. Dermatol. 2021, S0022-202X. [Google Scholar] [CrossRef]
  138. Muzumdar, S.; Koch, M.; Hiebert, H.; Bapst, A.; Gravina, A.; Bloch, W.; Beer, H.D.; Werner, S.; Schafer, M. Genetic activation of nrf2 reduces cutaneous symptoms in a murine model of netherton syndrome. Dis. Model. Mech. 2020, 13, 364–379. [Google Scholar] [CrossRef]
  139. Candi, E.; Melino, G.; Mei, G.; Tarcsa, E.; Chung, S.I.; Marekov, L.N.; Steinert, P.M. Biochemical, structural, and transglutaminase substrate properties of human loricrin, the major epidermal cornified cell envelope protein. J. Biol. Chem. 1995, 270, 26382–26390. [Google Scholar] [CrossRef] [Green Version]
  140. Steinert, P.M. The complexity and redundancy of epithelial barrier function. J. Cell Biol. 2000, 151, F5–F8. [Google Scholar] [CrossRef]
  141. Simon, M.; Jonca, N.; Guerrin, M.; Haftek, M.; Bernard, D.; Caubet, C.; Egelrud, T.; Schmidt, R.; Serre, G. Refined characterization of corneodesmosin proteolysis during terminal differentiation of human epidermis and its relationship to desquamation. J. Biol. Chem. 2001, 276, 20292–20299. [Google Scholar] [CrossRef] [Green Version]
  142. Eckert, R.L.; Yaffe, M.B.; Crish, J.F.; Murthy, S.; Rorke, E.A.; Welter, J.F. Involucrin--structure and role in envelope assembly. J. Investig. Dermatol. 1993, 100, 613–617. [Google Scholar] [CrossRef] [Green Version]
  143. Nemes, Z.; Steinert, P.M. Bricks and mortar of the epidermal barrier. Exp. Mol. Med. 1999, 31, 5–19. [Google Scholar] [CrossRef]
  144. Attar, P.S.; Wertz, P.W.; McArthur, M.; Imakado, S.; Bickenbach, J.R.; Roop, D.R. Inhibition of retinoid signaling in transgenic mice alters lipid processing and disrupts epidermal barrier function. Mol. Endocrinol. 1997, 11, 792–800. [Google Scholar] [CrossRef]
  145. Bickenbach, J.R.; Greer, J.M.; Bundman, D.S.; Rothnagel, J.A.; Roop, D.R. Loricrin expression is coordinated with other epidermal proteins and the appearance of lipid lamellar granules in development. J. Investig. Dermatol. 1995, 104, 405–410. [Google Scholar] [CrossRef] [Green Version]
  146. Matzinger, P. The danger model: A renewed sense of self. Science 2002, 296, 301–305. [Google Scholar] [CrossRef] [Green Version]
  147. Diven, D.G.; Dozier, S.E.; Meyer, D.J.; Smith, E.B. Bacteriology of inflamed and uninflamed epidermal inclusion cysts. Arch. Dermatol. 1998, 134, 49–51. [Google Scholar] [CrossRef] [Green Version]
  148. Dalziel, K.; Dykes, P.J.; Marks, R. Inflammation due to intra-cutaneous implantation of stratum corneum. Br. J. Exp. Pathol. 1984, 65, 107–115. [Google Scholar]
  149. Gahring, L.C.; Buckley, A.; Daynes, R.A. Presence of epidermal-derived thymocyte activating factor/interleukin 1 in normal human stratum corneum. J. Clin. Investig. 1985, 76, 1585–1591. [Google Scholar] [CrossRef] [Green Version]
  150. Kezic, S.; O’Regan, G.M.; Lutter, R.; Jakasa, I.; Koster, E.S.; Saunders, S.; Caspers, P.; Kemperman, P.M.; Puppels, G.J.; Sandilands, A.; et al. Filaggrin loss-of-function mutations are associated with enhanced expression of IL-1 cytokines in the stratum corneum of patients with atopic dermatitis and in a murine model of filaggrin deficiency. J. Allergy Clin. Immunol. 2012, 129, 1031–1039.e1. [Google Scholar] [CrossRef] [Green Version]
  151. O’Shaughnessy, R.F.; Choudhary, I.; Harper, J.I. Interleukin-1 alpha blockade prevents hyperkeratosis in an in vitro model of lamellar ichthyosis. Hum. Mol. Genet 2010, 19, 2594–2605. [Google Scholar] [CrossRef]
  152. Barland, C.O.; Zettersten, E.; Brown, B.S.; Ye, J.; Elias, P.M.; Ghadially, R. Imiquimod-induced interleukin-1 alpha stimulation improves barrier homeostasis in aged murine epidermis. J. Investig. Dermatol. 2004, 122, 330–336. [Google Scholar] [CrossRef] [Green Version]
  153. Jung, Y.J.; Jung, M.; Kim, M.; Hong, S.P.; Choi, E.H. IL-1α stimulation restores epidermal permeability and antimicrobial barriers compromised by topical tacrolimus. J. Investig. Dermatol. 2011, 131, 698–705. [Google Scholar] [CrossRef] [Green Version]
  154. Dinarello, C.A. Overview of the IL-1 family in innate inflammation and acquired immunity. Immunol. Rev. 2018, 281, 8–27. [Google Scholar] [CrossRef]
  155. Werman, A.; Werman-Venkert, R.; White, R.; Lee, J.K.; Werman, B.; Krelin, Y.; Voronov, E.; Dinarello, C.A.; Apte, R.N. The precursor form of IL-1alpha is an intracrine proinflammatory activator of transcription. Proc. Natl. Acad. Sci. USA 2004, 101, 2434–2439. [Google Scholar] [CrossRef] [Green Version]
  156. Lee, R.T.; Briggs, W.H.; Cheng, G.C.; Rossiter, H.B.; Libby, P.; Kupper, T. Mechanical deformation promotes secretion of IL-1 alpha and IL-1 receptor antagonist. J. Immunol. 1997, 159, 5084–5088. [Google Scholar]
  157. Nakae, S.; Naruse-Nakajima, C.; Sudo, K.; Horai, R.; Asano, M.; Iwakura, Y. IL-1 alpha, but not IL-1 beta, is required for contact-allergen-specific T cell activation during the sensitization phase in contact hypersensitivity. Int. Immunol. 2001, 13, 1471–1478. [Google Scholar] [CrossRef] [Green Version]
  158. Archer, N.K.; Jo, J.H.; Lee, S.K.; Kim, D.; Smith, B.; Ortines, R.V.; Wang, Y.; Marchitto, M.C.; Ravipati, A.; Cai, S.S.; et al. Injury, dysbiosis, and filaggrin deficiency drive skin inflammation through keratinocyte IL-1alpha release. J. Allergy Clin. Immunol. 2019, 143, 1426–1443.e6. [Google Scholar] [CrossRef] [Green Version]
  159. Ogawa, T.; Ishitsuka, Y.; Inoue, S.; Nakamura, Y.; Saito, A.; Okiyama, N.; Fujisawa, Y.; Furuta, J.; Watanabe, R.; Fujimoto, M. Nuclear Factor Erythroid 2-Related Factor 2 (Nrf2) Regulates Epidermal Keratinization under Psoriatic Skin Inflammation. Am. J. Pathol. 2020, 190, 577–585. [Google Scholar] [CrossRef]
  160. Bieber, T. Atopic dermatitis: An expanding therapeutic pipeline for a complex disease. Nat. Rev. Drug Discov. 2022, 21, 21–40. [Google Scholar] [CrossRef] [PubMed]
  161. Sevilla, L.M.; Nachat, R.; Groot, K.R.; Klement, J.F.; Uitto, J.; Djian, P.; Maatta, A.; Watt, F.M. Mice deficient in involucrin, envoplakin, and periplakin have a defective epidermal barrier. J. Cell Biol. 2007, 179, 1599–1612. [Google Scholar] [CrossRef] [Green Version]
  162. Jonca, N.; Guerrin, M.; Hadjiolova, K.; Caubet, C.; Gallinaro, H.; Simon, M.; Serre, G. Corneodesmosin, a component of epidermal corneocyte desmosomes, displays homophilic adhesive properties. J. Biol. Chem. 2002, 277, 5024–5029. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Simon, M.; Montezin, M.; Guerrin, M.; Durieux, J.J.; Serre, G. Characterization and purification of human corneodesmosin, an epidermal basic glycoprotein associated with corneocyte-specific modified desmosomes. J. Biol. Chem. 1997, 272, 31770–31776. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Serre, G.; Mils, V.; Haftek, M.; Vincent, C.; Croute, F.; Reano, A.; Ouhayoun, J.P.; Bettinger, S.; Soleilhavoup, J.P. Identification of late differentiation antigens of human cornified epithelia, expressed in re-organized desmosomes and bound to cross-linked envelope. J. Investig. Dermatol. 1991, 97, 1061–1072. [Google Scholar] [CrossRef] [Green Version]
  165. Haftek, M.; Serre, G.; Mils, V.; Thivolet, J. Immunocytochemical evidence for a possible role of cross-linked keratinocyte envelopes in stratum corneum cohesion. J. Histochem. Cytochem. 1991, 39, 1531–1538. [Google Scholar] [CrossRef] [Green Version]
  166. Rubartelli, A.; Lotze, M.T. Inside, outside, upside down: Damage-associated molecular-pattern molecules (DAMPs) and redox. Trends Immunol. 2007, 28, 429–436. [Google Scholar] [CrossRef]
  167. Doebel, T.; Voisin, B.; Nagao, K. Langerhans Cells—The Macrophage in Dendritic Cell Clothing. Trends Immunol. 2017, 38, 817–828. [Google Scholar] [CrossRef]
  168. Kaplan, D.H. Ontogeny and function of murine epidermal Langerhans cells. Nat. Immunol. 2017, 18, 1068–1075. [Google Scholar] [CrossRef]
  169. Steinman, R.M.; Nussenzweig, M.C. Avoiding horror autotoxicus: The importance of dendritic cells in peripheral T cell tolerance. Proc. Natl. Acad. Sci. USA 2002, 99, 351–358. [Google Scholar] [CrossRef] [Green Version]
  170. Zimmerli, S.C.; Hauser, C. Langerhans cells and lymph node dendritic cells express the tight junction component claudin-1. J. Investig. Dermatol. 2007, 127, 2381–2390. [Google Scholar] [CrossRef] [Green Version]
  171. Gomez de Aguero, M.; Vocanson, M.; Hacini-Rachinel, F.; Taillardet, M.; Sparwasser, T.; Kissenpfennig, A.; Malissen, B.; Kaiserlian, D.; Dubois, B. Langerhans cells protect from allergic contact dermatitis in mice by tolerizing CD8(+) T cells and activating Foxp3(+) regulatory T cells. J. Clin. Investig. 2012, 122, 1700–1711. [Google Scholar] [CrossRef]
  172. Kaplan, D.H.; Jenison, M.C.; Saeland, S.; Shlomchik, W.D.; Shlomchik, M.J. Epidermal langerhans cell-deficient mice develop enhanced contact hypersensitivity. Immunity 2005, 23, 611–620. [Google Scholar] [CrossRef] [Green Version]
  173. Zhang, S.; Edwards, T.N.; Chaudhri, V.K.; Wu, J.; Cohen, J.A.; Hirai, T.; Rittenhouse, N.; Schmitz, E.G.; Zhou, P.Y.; McNeil, B.D.; et al. Nonpeptidergic neurons suppress mast cells via glutamate to maintain skin homeostasis. Cell 2021, 184, 2151–2166.e16. [Google Scholar] [CrossRef]
  174. Lee, H.J.; Kim, T.G.; Kim, S.H.; Park, J.Y.; Lee, M.; Lee, J.W.; Lee, S.H.; Lee, M.G. Epidermal Barrier Function Is Impaired in Langerhans Cell-Depleted Mice. J. Investig. Dermatol. 2019, 139, 1182–1185. [Google Scholar] [CrossRef]
  175. Capucha, T.; Mizraji, G.; Segev, H.; Blecher-Gonen, R.; Winter, D.; Khalaileh, A.; Tabib, Y.; Attal, T.; Nassar, M.; Zelentsova, K.; et al. Distinct Murine Mucosal Langerhans Cell Subsets Develop from Pre-dendritic Cells and Monocytes. Immunity 2015, 43, 369–381. [Google Scholar] [CrossRef] [Green Version]
  176. Allen, T.D.; Potten, C.S. Fine-structural identification and organization of the epidermal proliferative unit. J. Cell Sci. 1974, 15, 291–319. [Google Scholar] [CrossRef]
  177. Bauer, J.; Bahmer, F.A.; Worl, J.; Neuhuber, W.; Schuler, G.; Fartasch, M. A strikingly constant ratio exists between Langerhans cells and other epidermal cells in human skin. A stereologic study using the optical disector method and the confocal laser scanning microscope. J. Investig. Dermatol. 2001, 116, 313–318. [Google Scholar] [CrossRef]
  178. Park, S.; Matte-Martone, C.; Gonzalez, D.G.; Lathrop, E.A.; May, D.P.; Pineda, C.M.; Moore, J.L.; Boucher, J.D.; Marsh, E.; Schmitter-Sanchez, A.; et al. Skin-resident immune cells actively coordinate their distribution with epidermal cells during homeostasis. Nat. Cell Biol. 2021, 23, 476–484. [Google Scholar] [CrossRef]
  179. Ramirez, K.; Witherden, D.A.; Havran, W.L. All hands on DE(T)C: Epithelial-resident gammadelta T cells respond to tissue injury. Cell. Immunol. 2015, 296, 57–61. [Google Scholar] [CrossRef] [Green Version]
  180. Mohammed, J.; Beura, L.K.; Bobr, A.; Astry, B.; Chicoine, B.; Kashem, S.W.; Welty, N.E.; Igyarto, B.Z.; Wijeyesinghe, S.; Thompson, E.A.; et al. Stromal cells control the epithelial residence of DCs and memory T cells by regulated activation of TGF-beta. Nat. Immunol. 2016, 17, 414–421. [Google Scholar] [CrossRef]
  181. Schon, M.P.; Schon, M.; Parker, C.M.; Williams, I.R. Dendritic epidermal T cells (DETC) are diminished in integrin alphaE(CD103)-deficient mice. J. Investig. Dermatol. 2002, 119, 190–193. [Google Scholar] [CrossRef] [Green Version]
  182. Lefrancois, L.; Barrett, T.A.; Havran, W.L.; Puddington, L. Developmental expression of the alpha IEL beta 7 integrin on T cell receptor gamma delta and T cell receptor alpha beta T cells. Eur. J. Immunol. 1994, 24, 635–640. [Google Scholar] [CrossRef] [PubMed]
  183. De La Cruz Diaz, J.S.; Hirai, T.; Anh-Thu Nguyen, B.; Zenke, Y.; Yang, Y.; Li, H.; Nishimura, S.; Kaplan, D.H. TNF-alpha and IL-1beta Do Not Induce Langerhans Cell Migration by Inhibiting TGFbeta Activation. JID Innov. 2021, 1, 100028. [Google Scholar] [CrossRef] [PubMed]
  184. Ginhoux, F.; Tacke, F.; Angeli, V.; Bogunovic, M.; Loubeau, M.; Dai, X.M.; Stanley, E.R.; Randolph, G.J.; Merad, M. Langerhans cells arise from monocytes in vivo. Nat. Immunol. 2006, 7, 265–273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Merad, M.; Manz, M.G.; Karsunky, H.; Wagers, A.; Peters, W.; Charo, I.; Weissman, I.L.; Cyster, J.G.; Engleman, E.G. Langerhans cells renew in the skin throughout life under steady-state conditions. Nat. Immunol. 2002, 3, 1135–1141. [Google Scholar] [CrossRef] [Green Version]
  186. Bobr, A.; Igyarto, B.Z.; Haley, K.M.; Li, M.O.; Flavell, R.A.; Kaplan, D.H. Autocrine/paracrine TGF-beta1 inhibits Langerhans cell migration. Proc. Natl. Acad. Sci. USA 2012, 109, 10492–10497. [Google Scholar] [CrossRef] [Green Version]
  187. Gaiser, M.R.; Lammermann, T.; Feng, X.; Igyarto, B.Z.; Kaplan, D.H.; Tessarollo, L.; Germain, R.N.; Udey, M.C. Cancer-associated epithelial cell adhesion molecule (EpCAM.; CD326) enables epidermal Langerhans cell motility and migration in vivo. Proc. Natl. Acad. Sci. USA 2012, 109, E889–E897. [Google Scholar] [CrossRef] [Green Version]
  188. 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]
  189. Jiang, A.; Bloom, O.; Ono, S.; Cui, W.; Unternaehrer, J.; Jiang, S.; Whitney, J.A.; Connolly, J.; Banchereau, J.; Mellman, I. Disruption of E-cadherin-mediated adhesion induces a functionally distinct pathway of dendritic cell maturation. Immunity 2007, 27, 610–624. [Google Scholar] [CrossRef] [Green Version]
  190. Shklovskaya, E.; O’Sullivan, B.J.; Ng, L.G.; Roediger, B.; Thomas, R.; Weninger, W.; Fazekas de St Groth, B. Langerhans cells are precommitted to immune tolerance induction. Proc. Natl. Acad. Sci. USA 2011, 108, 18049–18054. [Google Scholar] [CrossRef] [Green Version]
  191. Haley, K.; Igyarto, B.Z.; Ortner, D.; Bobr, A.; Kashem, S.; Schenten, D.; Kaplan, D.H. Langerhans cells require MyD88-dependent signals for Candida albicans response but not for contact hypersensitivity or migration. J. Immunol. 2012, 188, 4334–4339. [Google Scholar] [CrossRef] [Green Version]
  192. Baratin, M.; Foray, C.; Demaria, O.; Habbeddine, M.; Pollet, E.; Maurizio, J.; Verthuy, C.; Davanture, S.; Azukizawa, H.; Flores-Langarica, A.; et al. Homeostatic NF-kappaB Signaling in Steady-State Migratory Dendritic Cells Regulates Immune Homeostasis and Tolerance. Immunity 2015, 42, 627–639. [Google Scholar] [CrossRef] [Green Version]
  193. Kissenpfennig, A.; Henri, S.; Dubois, B.; Laplace-Builhe, C.; Perrin, P.; Romani, N.; Tripp, C.H.; Douillard, P.; Leserman, L.; Kaiserlian, D.; et al. Dynamics and function of Langerhans cells in vivo: Dermal dendritic cells colonize lymph node areas distinct from slower migrating Langerhans cells. Immunity 2005, 22, 643–654. [Google Scholar] [CrossRef] [Green Version]
  194. Watanabe, H.; Bannai, S. Induction of cystine transport activity in mouse peritoneal macrophages. J. Exp. Med. 1987, 165, 628–640. [Google Scholar] [CrossRef]
  195. Sasaki, H.; Sato, H.; Kuriyama-Matsumura, K.; Sato, K.; Maebara, K.; Wang, H.; Tamba, M.; Itoh, K.; Yamamoto, M.; Bannai, S. Electrophile response element-mediated induction of the cystine/glutamate exchange transporter gene expression. J. Biol. Chem. 2002, 277, 44765–44771. [Google Scholar] [CrossRef] [Green Version]
  196. Gosselin, D.; Skola, D.; Coufal, N.G.; Holtman, I.R.; Schlachetzki, J.C.M.; Sajti, E.; Jaeger, B.N.; O’Connor, C.; Fitzpatrick, C.; Pasillas, M.P.; et al. An environment-dependent transcriptional network specifies human microglia identity. Science 2017, 356, eaal3222. [Google Scholar] [CrossRef] [Green Version]
  197. Guilliams, M.; Svedberg, F.R. Does tissue imprinting restrict macrophage plasticity? Nat. Immunol. 2021, 22, 118–127. [Google Scholar] [CrossRef]
  198. Su, Q.; Igyarto, B.Z. Keratinocytes Share Gene Expression Fingerprint with Epidermal Langerhans Cells via mRNA Transfer. J. Investig. Dermatol. 2019, 139, 2313–2323.e8. [Google Scholar] [CrossRef]
  199. Acharya, M.; Mukhopadhyay, S.; Paidassi, H.; Jamil, T.; Chow, C.; Kissler, S.; Stuart, L.M.; Hynes, R.O.; Lacy-Hulbert, A. alphav Integrin expression by DCs is required for Th17 cell differentiation and development of experimental autoimmune encephalomyelitis in mice. J. Clin. Investig. 2010, 120, 4445–4452. [Google Scholar] [CrossRef]
  200. Worthington, J.J.; Kelly, A.; Smedley, C.; Bauche, D.; Campbell, S.; Marie, J.C.; Travis, M.A. Integrin alphavbeta8-Mediated TGF-beta Activation by Effector Regulatory T Cells Is Essential for Suppression of T-Cell-Mediated Inflammation. Immunity 2015, 42, 903–915. [Google Scholar] [CrossRef] [Green Version]
  201. Nishibu, A.; Ward, B.R.; Jester, J.V.; Ploegh, H.L.; Boes, M.; Takashima, A. Behavioral responses of epidermal Langerhans cells in situ to local pathological stimuli. J. Investig. Dermatol. 2006, 126, 787–796. [Google Scholar] [CrossRef] [Green Version]
  202. Shelley, W.B.; Juhlin, L. Langerhans cells form a reticuloepithelial trap for external contact antigens. Nature 1976, 261, 46–47. [Google Scholar] [CrossRef]
  203. Mani, V.; Bromley, S.K.; Aijo, T.; Mora-Buch, R.; Carrizosa, E.; Warner, R.D.; Hamze, M.; Sen, D.R.; Chasse, A.Y.; Lorant, A.; et al. Migratory DCs activate TGF-beta to precondition naive CD8(+) T cells for tissue-resident memory fate. Science 2019, 366, eaav5728. [Google Scholar] [CrossRef]
  204. Bynoe, M.S.; Evans, J.T.; Viret, C.; Janeway, C.A., Jr. Epicutaneous immunization with autoantigenic peptides induces T suppressor cells that prevent experimental allergic encephalomyelitis. Immunity 2003, 19, 317–328. [Google Scholar] [CrossRef] [Green Version]
  205. Walczak, A.; Siger, M.; Ciach, A.; Szczepanik, M.; Selmaj, K. Transdermal application of myelin peptides in multiple sclerosis treatment. JAMA Neurol. 2013, 70, 1105–1109. [Google Scholar] [CrossRef]
  206. Tordesillas, L.; Lozano-Ojalvo, D.; Dunkin, D.; Mondoulet, L.; Agudo, J.; Merad, M.; Sampson, H.A.; Berin, M.C. PDL2(+) CD11b(+) dermal dendritic cells capture topical antigen through hair follicles to prime LAP(+) Tregs. Nat. Commun. 2018, 9, 5238. [Google Scholar] [CrossRef]
  207. Tordesillas, L.; Mondoulet, L.; Blazquez, A.B.; Benhamou, P.H.; Sampson, H.A.; Berin, M.C. Epicutaneous immunotherapy induces gastrointestinal LAP(+) regulatory T cells and prevents food-induced anaphylaxis. J. Allergy Clin. Immunol. 2017, 139, 189–201.e4. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Loricrin (LOR) may directly affect Langerhans cell (LC) differentiation (maturation) and imprint cutaneous immune effector functions. This regulation requires direct interaction between LCs and differentiating keratinocytes (KCs) in the stratum granulosum (SG). LOR anchors differentiate keratins K1/K10 and corneodesmosomes, establishing the desmosome (DM)-keratin (KIF) scaffold in the stratum corneum (SC). LCs are anchored to KCs through integrin αVβ6 expressed on the KC plasma membrane, and activate endogenous transforming growth factor-beta (TGF-β) signaling for epidermal retention. The prodomain (proTGF-β) “fastens” free TGF-β via disulfide (–S–S–) cross-linkages in the bowtie and straitjacket regions. The arm domain anchors keratinocytes via αVβ6 integrin and negatively regulates TGF-β in the extracellular matrix. Therefore, successful cornification may involve disulfide (–S–S–) cross-linkages of the junctional component, possibly inactivating TGF-β permanently and promoting LC maturation. Because LOR promotes the structural maturation of corneodesmosomes (compactization of corneocytes), a delay in the maturation of junctional components would cause the retention of adheren junctions (AJs)/tight junctions (TJs) to the more apical end. The delay in the structural maturation of the LOR-knockout (LKO) epidermis may be advantageous for LCs to remain active (mobile and endocytotic) in situ. The immature structure of the LKO epidermis may imprint adaptive immune responses through more increased TGF-β bioavailability than wild-type (WT) epidermis, causing immunological tolerance (or immune privilege). CE—cornified cell envelope; DC—dendritic cell; BM—basal membrane; SS—stratum spinosum; EPU—epidermal differentiation unit; EDU—epidermal differentiation unit.
Figure 1. Loricrin (LOR) may directly affect Langerhans cell (LC) differentiation (maturation) and imprint cutaneous immune effector functions. This regulation requires direct interaction between LCs and differentiating keratinocytes (KCs) in the stratum granulosum (SG). LOR anchors differentiate keratins K1/K10 and corneodesmosomes, establishing the desmosome (DM)-keratin (KIF) scaffold in the stratum corneum (SC). LCs are anchored to KCs through integrin αVβ6 expressed on the KC plasma membrane, and activate endogenous transforming growth factor-beta (TGF-β) signaling for epidermal retention. The prodomain (proTGF-β) “fastens” free TGF-β via disulfide (–S–S–) cross-linkages in the bowtie and straitjacket regions. The arm domain anchors keratinocytes via αVβ6 integrin and negatively regulates TGF-β in the extracellular matrix. Therefore, successful cornification may involve disulfide (–S–S–) cross-linkages of the junctional component, possibly inactivating TGF-β permanently and promoting LC maturation. Because LOR promotes the structural maturation of corneodesmosomes (compactization of corneocytes), a delay in the maturation of junctional components would cause the retention of adheren junctions (AJs)/tight junctions (TJs) to the more apical end. The delay in the structural maturation of the LOR-knockout (LKO) epidermis may be advantageous for LCs to remain active (mobile and endocytotic) in situ. The immature structure of the LKO epidermis may imprint adaptive immune responses through more increased TGF-β bioavailability than wild-type (WT) epidermis, causing immunological tolerance (or immune privilege). CE—cornified cell envelope; DC—dendritic cell; BM—basal membrane; SS—stratum spinosum; EPU—epidermal differentiation unit; EDU—epidermal differentiation unit.
Biomolecules 12 00673 g001
Table 1. ARCI.
Table 1. ARCI.
ARCI #GeneFunctionCategoryReference #
1TGM1Catalyzing ε-(γ-glutamyl) lysine isopeptide bondsiii[95]
2ALOX12BLinoleic acid metabolismi[93]
3ALOXE3Linoleic acid metabolismi[93]
4ABCA12Lipid transport and ceramide linoleic ester formationii[94]
5CYP4F22Catalyzing ultra-long-chain fatty acidsi[88]
6NIPAL4Mg2+ transporti[92]
7N/DN/AN/AN/A
8LIPNLipasei[91]
9CERS3C24-ceramides synthesisi[89]
10PNPLA1Linoleic acid estelificationi[90]
11ST14Filaggrin processingiv[97]
12CASP14Filaggrin processingiv[96]
13SDRC9C7Ceramide linoleic ester formationi[86]
14SULT2B1Sulfoconjugation of neutral steroids and sterolsi[87]
Abbreviations: TGM1—transglutaminase 1; ALOX12B—arachidonate 12-lipoxygenase; ALOXE3—arachidonate lipoxygenase 3; ABCA12—ATP-binding cassette subfamily A member 12; CYP4F22—cytochrome P450, family 4, subfamily F, polypeptide 22; NIPAL4—NIPA-like domain-containing 4; LIPN—lipase family member N; CERS3—ceramide synthase 3; PNPLA1—patatin-like phospholipase domain-containing 1; ST14—suppressor of tumorigenicity 14 protein; CASP14—caspase 14; SDRC9C7—short-chain dehydrogenase/reductase family 9C member 7; SULT2B1—sulfotransferase family 2B member 1. N/A: not applicable
Table 2. Desmosomal Defects Accompanying Allergic Manifestations.
Table 2. Desmosomal Defects Accompanying Allergic Manifestations.
Clinical
Nomenclatures
GeneFunctionReference #
Netherton syndrome
ichthyosis linearis circumflexia
SPINK5Inhibiting serin proteases that degrade corneodesmosomes[79]
Peeling skin syndrome-1/
Peeling skin disease
CDSNMaintaining corneodesmosomal adhesion[98]
Skin dermatitis, multiple severe allergies, and metabolic wasting (SAM) syndromeDSG1Maintaining the desmosomal-keratin scaffold in the differentiating layers[80]
Abbreviations: SPINK5—serine protease inhibitor of Kasal-type 5; CDSN—corneodesmosin; DSG1—desmoglein 1.
Table 3. Ichthyosiform Dermatosis Caused by LOR Mislocalization.
Table 3. Ichthyosiform Dermatosis Caused by LOR Mislocalization.
Clinical
Nomenclature
GeneFunctionReference #
Vohwinkel syndrome with ichthyosisLORLocalizing to the cell periphery and replacing the plasma cell membrane
NLS pertubing cornification
[84,103]
Abbreviations: LOR—loricrin; NLS—nuclear localization signal.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ishitsuka, Y.; Roop, D.R. Loricrin at the Boundary between Inside and Outside. Biomolecules 2022, 12, 673. https://doi.org/10.3390/biom12050673

AMA Style

Ishitsuka Y, Roop DR. Loricrin at the Boundary between Inside and Outside. Biomolecules. 2022; 12(5):673. https://doi.org/10.3390/biom12050673

Chicago/Turabian Style

Ishitsuka, Yosuke, and Dennis R. Roop. 2022. "Loricrin at the Boundary between Inside and Outside" Biomolecules 12, no. 5: 673. https://doi.org/10.3390/biom12050673

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