<<

Brief Exposure of Skin to Near-Infrared Laser Modulates Mast Cell Function and Augments the Immune Response

This information is current as Yoshifumi Kimizuka, Wataru Katagiri, Joseph J. Locascio, of September 27, 2021. Ayako Shigeta, Yuri Sasaki, Mai Shibata, Kaitlyn Morse, Ruxandra F. Sîrbulescu, Mizuki Miyatake, Patrick Reeves, Makoto Suematsu, Jeffrey Gelfand, Timothy Brauns, Mark C. Poznansky, Kosuke Tsukada and Satoshi Kashiwagi

J Immunol published online 12 November 2018 Downloaded from http://www.jimmunol.org/content/early/2018/11/10/jimmun ol.1701687

Supplementary http://www.jimmunol.org/content/suppl/2018/11/10/jimmunol.170168 http://www.jimmunol.org/ Material 7.DCSupplemental

Why The JI? Submit online.

• Rapid Reviews! 30 days* from submission to initial decision

• No Triage! Every submission reviewed by practicing scientists by guest on September 27, 2021

• Fast Publication! 4 weeks from acceptance to publication

*average

Subscription Information about subscribing to The Journal of Immunology is online at: http://jimmunol.org/subscription Permissions Submit copyright permission requests at: http://www.aai.org/About/Publications/JI/copyright.html Email Alerts Receive free email-alerts when new articles cite this article. Sign up at: http://jimmunol.org/alerts

The Journal of Immunology is published twice each month by The American Association of Immunologists, Inc., 1451 Rockville Pike, Suite 650, Rockville, MD 20852 Copyright © 2018 by The American Association of Immunologists, Inc. All rights reserved. Print ISSN: 0022-1767 Online ISSN: 1550-6606. Published November 12, 2018, doi:10.4049/jimmunol.1701687 The Journal of Immunology

Brief Exposure of Skin to Near-Infrared Laser Modulates Mast Cell Function and Augments the Immune Response

Yoshifumi Kimizuka,* Wataru Katagiri,†,‡,x Joseph J. Locascio,{ Ayako Shigeta,* Yuri Sasaki,* Mai Shibata,* Kaitlyn Morse,* Ruxandra F. Sıˆrbulescu,* Mizuki Miyatake,‖ Patrick Reeves,* Makoto Suematsu,# Jeffrey Gelfand,* Timothy Brauns,* Mark C. Poznansky,*,1 Kosuke Tsukada,‡,‖,1 and Satoshi Kashiwagi*,†

The treatment of skin with a low-power continuous-wave (CW) near-infrared (NIR) laser prior to vaccination is an emerging strat- egy to augment the immune response to intradermal vaccine, potentially substituting for chemical adjuvant, which has been linked to adverse effects of vaccines. This approach proved to be low cost, simple, small, and readily translatable compared with the

previously explored pulsed-wave medical lasers. However, little is known on the mode of laser–tissue interaction eliciting the Downloaded from adjuvant effect. In this study, we sought to identify the pathways leading to the immunological events by examining the alteration of responses resulting from genetic ablation of innate subsets including mast cells and specific dendritic cell populations in an established model of intradermal vaccination and analyzing functional changes of skin microcirculation upon the CW NIR laser treatment in mice. We found that a CW NIR laser transiently stimulates mast cells via generation of reactive oxygen species, establishes an immunostimulatory milieu in the exposed tissue, and provides migration cues for dermal CD103+ dendritic cells

without inducing prolonged inflammation, ultimately augmenting the adaptive immune response. These results indicate that http://www.jimmunol.org/ use of an NIR laser with distinct wavelength and power is a safe and effective tool to reproducibly modulate innate programs in skin. These mechanistic findings would accelerate the clinical translation of this technology and warrant further explorations into the broader application of NIR lasers to the treatment of immune-related skin diseases. The Journal of Immunology,2018, 201: 000–000.

kin-based vaccination has been demonstrated to permit little protective T cell response, conferring insufficient protection to dose sparing of the vaccine Ag and improve long-term vulnerable populations, including children and the elderly (4–6). The protection conferred by vaccination, thus showing clear shortcomings of the current inactivated vaccine strategies would

S by guest on September 27, 2021 advantages over the standard i.m. injection (1–3). However, be improved by use of immunologic adjuvants, as incorporation of modern inactivated vaccines used for intradermal (ID) influenza an adjuvant into vaccine formulations is an established strategy to vaccination, for instance, are still poorly immunogenic and induce enhance immunogenicity of a vaccine Ag and induce long-term

*Vaccine and Immunotherapy Center, Division of Infectious Diseases, Department of AI060354-10). The Massachusetts General Hospital Department of Pathology Flow Medicine, Massachusetts General Hospital, Charlestown, MA 02129; †Gordon Cen- and Image Cytometry Research Core obtained support from the NIH Shared Instru- ter for Medical Imaging, Department of Radiology, Massachusetts General Hospital, mentation Program through Grants 1S10OD012027-01A1, 1S10OD016372-01, Charlestown, MA 02129; ‡Graduate School of Fundamental Science and Technology, 1S10RR020936-01, and 1S10RR023440-01A1. Keio University, Yokohama, Kanagawa 223-8522, Japan; xSchool of Engineering Y.K. performed most of the experiments; Y.K., A.S., Y.S., M. Shibata, K.M., and S.K. Sciences in Chemistry, Biotechnology and Health, Royal Institute of Technology, { performed the vaccination studies; Y.K., W.K., M.M., K.T., and P.R. contributed to the 14152 Huddinge, Sweden; Alzheimer’s Disease Research Center, Department of ‖ analysis and experimental design related to the reactive oxygen species (ROS) study; Y.K., Neurology, Massachusetts General Hospital, Boston, MA 02114; Faculty of Sci- M. Suematsu, K.T., and S.K. contributed to the expression study; Y.K., W.K., R.F.S., ence and Technology, Keio University, Yokohama, Kanagawa 223-8522, Japan; K.T., and S.K. contributed to the analysis and experimental design related to the skin and #Department of Biochemistry, Keio University School of Medicine, Shinjuku- histology study; Y.K., W.K., M.M., and K.T. performed the in vitro laser study using bone ku, Tokyo 160-0016, Japan marrow–derived mast cells and keratinocytes; S.K. designed the study; J.J.L. performed 1M.C.P. and K.T. contributed equally to this work. statistical analysis on Ab responses and quantitative PCR results and advised on statistical tests in this study; and Y.K., J.G., T.B., M.C.P., K.T., and S.K. wrote the manuscript. ORCIDs: 0000-0002-4026-7506 (Y.K.); 0000-0003-3439-1209 (J.J.L.); 0000-0003- 2369-9574 (A.S.); 0000-0002-3077-2832 (K.M.); 0000-0001-7905-1713 (R.F.S.); The funders had no role in study design, data collection and analysis, decision to 0000-0002-5387-6271 (M.M.); 0000-0001-8202-0723 (P.R.); 0000-0002-4118- publish, or preparation of the manuscript. 5676 (J.G.); 0000-0002-2161-551X (T.B.). Address correspondence and reprint requests to Dr. Satoshi Kashiwagi, Gordon Received for publication December 6, 2017. Accepted for publication October 15, Center for Medical Imaging, Department of Radiology, Massachusetts General 2018. Hospital, 149 13th Street, Charlestown, MA 02129. E-mail address: skashiwagi@mgh. harvard.edu This work was supported by funds from the National Institute of Allergy and Infec- tious Diseases/National Institutes of Health (NIH) under Award R01AI105131 (to The online version of this article contains supplemental material. S.K.), the Kanae Foundation for the Promotion of Medical Science (to Y.K.), the 2 2 Abbreviations used in this article: Batf3 / , Batf3-deficient; BM, bone marrow; Mochida Memorial Foundation for Medical and Pharmaceutical Research (to Y.K.), 2 2 BMMC, bone marrow–derived MC; Ccr7 / , Ccr7-deficient; cDC, classical DC; the Uehara Memorial Foundation (to Y.K.), the Japan Foundation for Pediatric Re- CW, continuous-wave; DC, dendritic cell; dLN, draining node; HAI, hemag- search (to Y.K.), the Vaccine and Immunotherapy Center Innovation Fund (to S.K.), a glutination inhibition; ID, intradermal(ly); IR, infrared; Lang+, langerin+; LN, lymph Keio University Special Grant-in-Aid for Innovative Collaborative Research Projects node; MC, mast cell; MGH, Massachusetts General Hospital; migDC, migratory DC; (to K.T.), Keio University Research Grants for Global Initiative Research Projects (to NAC, N-acetyl-L-cysteine; NIR, near-infrared; PW, pulsed-wave; ROS, reactive ox- K.T.), and the Japan Agency for Medical Research and Development Translational ygen species; WT, wild-type. Research Network Program (to K.T.). The Live Cell Imaging Facility, Karolinska Institutet, Sweden, is supported by grants from the Knut and Alice Wallenberg Ó Foundation. The Ragon Institute Imaging Core Facility is supported in part by the Copyright 2018 by The American Association of Immunologists, Inc. 0022-1767/18/$37.50 Harvard University Center for AIDS Research, an NIH-funded program (5 P30

www.jimmunol.org/cgi/doi/10.4049/jimmunol.1701687 2 NEAR-INFRARED LASER MODULATES MAST CELLS IN SKIN immunity and protective T cell responses (7). Unfortunately, most 013755), KitW-sh/KitW-sh (012861), and CD11c-EYFP (008829) mice from conventional chemical or biological adjuvants are inappropriate for The Jackson Laboratory and bred them at Massachusetts General Hospital use in the skin because of issues related to the potential to induce (MGH). All animal procedures were performed following the Public Health Service Policy on Humane Care of Laboratory Animals and ap- persistent local toxicity caused by overstimulated innate immunity in proved by the Institutional Animal Care and Use Committee of MGH. the confined space (8, 9). As a result, no adjuvant has been incor- porated into clinically approved ID vaccines. NIR laser treatment Recently, stimulating the skin with a physical parameter, namely We applied a Nd:YVO4 1064-nm laser emitting either CW or nanosecond laser light, prior to the administration of an ID vaccine has been shown PW at a repetition rate of 10 kHz (RMI Laser, Lafayette, CO) on the to augment the immunogenicity without applicable adverse effects, surface of depilated skin as previously described (19, 20). We used a non– 2 representing an emerging alternative approach to the conventional tissue damaging power density of 5 W/cm on a circular spot on the skin surface of ∼5 mm in diameter (0.2 cm2) for 1 min with a total dose of chemical or biological adjuvants (10). To date, four classes of lasers 300 J/cm2 as previously determined (19, 20). We monitored the skin have been used to enhance the efficacy of vaccines in preclinical temperature throughout the laser application using an infrared (IR) thermal mouse models (11). The application of a nonablative fractional laser imager (FLIR Systems). has been demonstrated to augment the immune response to a model Influenza vaccination models ID influenza vaccine via release of damage-associated molecular patterns from columns of coagulated (dead) tissue created after the We depilated the mouse back skin with commercial depilatory cream (Nair; treatment of skin with a fractional laser and subsequent recruitment Church & Dwight) 2 d before the NIR laser treatment and vaccination. We injected an inactivated influenza virus vaccine (A/PR/8/34, 1 mgin10ml of plasmacytoid dendritic cells (DCs) around the columns (12, 13). of saline; Charles River) ID in the center of the spot on the mouse back An ablative fractional laser has been similarly shown to activate skin within 5 min of the laser treatment. The vaccine mixed with Alum Downloaded from skin-resident Xcr1+ DCs and enhance the efficacy of prophylactic (diluted 1:1 v/v, Imject; Thermo Fisher Scientific) or c48/80 (3.2 ng/spot; and therapeutic tumor vaccines through the local inflammatory re- Millipore Sigma) was used for comparison as appropriate. We homo- sponse induced by dead keratinocytes around skin micropores (14). logously challenged mice intranasally with live influenza A/PR/8/34 virus at a dose of 2 3 105 50% egg infectious doses at day 28. We euthanized In contrast, brief exposures of skin with a low-power, visible pulsed- mice and obtained and samples for further analysis 4 d after wave (PW) laser were also found to augment the immune response challenge as established previously (18–20).

to a model ID vaccine via release of hsp70 from skin cells and http://www.jimmunol.org/ subsequent activation of skin-resident Langerhans cells (15) or dis- Antioxidant treatment ruption of cell–matrix interactions induced by disarray of dermal We examined the involvement of ROS in the effect of an NIR laser using a collagen fibers and resultant increase in the motility of APCs in the ROS scavenger. We pretreated mice with s.c. injections of 100 mg/kg 2 2 2 skin tissue (9, 16, 17). In contrast to the fractional lasers, this class of N-acetyl-L-cysteine (NAC; Millipore Sigma) at day 3, 2, 1, and 0 as appropriate (22) before the experiment. Control mice received vehicle visible PW laser induces no histological skin tissue damage (9), but only (PBS) s.c. its condensed energy in laser pulses induces the heat shock re- sponse (15) or ultrastructural change in the extracellular matrix (17). Anti-influenza Ab and hemagglutination inhibition titers We have shown that another distinct class of laser, a low-power We measured anti-influenza–specific IgG, IgG1, and IgG2c humoral re- continuous-wave (CW) near-infrared (NIR) invisible laser between sponses by ELISA as previously described (19). We coated Immulon plates by guest on September 27, 2021 1061and1301nm,significantlyenhances the immune response to a (Thermo Fisher Scientific) with 100 ng of the inactivated influenza virus model ID vaccine and improves survival in a lethal challenge murine and then added serially diluted mouse serum samples to the wells. We detected bound Igs with the secondary Ab (goat anti-mouse IgG [1:10,000; influenza model (18–20). In these studies, exposure of the skin to the Millipore Sigma], rat anti-mouse IgG1 [1:2000; SouthernBiotech], or goat CW NIR laser was found to induce no detectable tissue damage or anti-mouse IgG2c [1:4000; SouthernBiotech]). We designated a titer as the inflammation (18–20). The energy of a CW NIR laser, which is serum dilution corresponding to the inflection point of the plot of the OD delivered homogenously over the time period, is not high enough to versus dilution of serum. SRI International (Harrisonburg, VA) determined hemagglutination inhibition (HAI) titers in sera samples. Operators per- induce any heat shock response or ultrastructural change in the ex- formed all immunoassays in a blinded manner to control or experimental posed tissue (21). These observations indicate that the CW NIR laser groups. has the distinct mechanisms of action compared with the other three PW lasers. We have demonstrated that a CW NIR laser induces Assessment of T cell responses transient upregulation of a selective set of chemokines in skin and We harvested and immediately processed splenocytes 4 d after challenge as migration of specific subsets of skin-resident DCs (19, 20). However, previously described (18–20). Briefly, we incubated splenocyte prepara- 6 the precise details of CW NIR laser–skin tissue and cell–cell inter- tions containing 1 3 10 cells with or without the inactive influenza at a concentration of 1 mg/ml for 60 h for determination of cytokine release in a actions leading to the augmentation of the immune response are round-bottom 96-well plate. We collected splenocyte culture supernatants largely unknown. and determined the amounts of IFN-g or IL-4 using the DuoSet ELISA Kit In this study, we demonstrate that a CW NIR laser induces a (R&D Systems) following manufacturer’s instructions. Operators per- unique immunologic cascade in skin-stimulating mast cells formed all immunoassays in a blinded manner to control or experimental (MCs) via reactive oxygen species (ROS) generation and cre- groups. ates immunostimulatory milieu for DCs, thus augmenting the Influenza virus challenge study adaptive immune response. These findings not only contribute to We challenged immunized mice intranasally with live influenza A/PR/8/34 the optimization of a CW NIR laser for immunologic adjuvant virus at a dose of 1.5 3 106 50% egg infectious doses, which is equivalent but open a pathway to broader applications of this technology to to 150 3 50% mouse lethal dose, in 30 ml of saline 28 d after vaccination treat immune-related conditions. as previously described (18–20). We monitored survival and body weight for 15 d postchallenge. We considered mice showing a hunched posture, ruffled fur, or .20% body weight loss or mice that were not eating or Materials and Methods drinking to have reached the experimental endpoint. Operators performed Animals monitoring in a blinded manner to control or experimental groups. We purchased 8-wk-old female C57BL/6J mice (stock number: 000664) FITC painting assay from The Jackson Laboratory (Bar Harbor, ME). We acclimated all ani- mals for at least 2 wk prior to the experiments. We purchased breeding We shaved and depilated mice 2 d before the assay. We painted mice on four pairs of Ccr7-deficient (Ccr72/2; 006621), Batf3-deficient (Batf32/2; spots of the flank back skin (∼5 mm in diameter per spot) with 10 mlofa The Journal of Immunology 3

0.5% FITC solution (Isomer I; Millipore Sigma) 4 h prior to the assay as Detection of ROS in cultured murine BMMCs and previously described (20, 23). The FITC solution was prepared in acetone: human keratinocytes dibutyl phthalate (1:1, vol/vol; Millipore Sigma). We performed the NIR laser treatment followed by vaccination with OVA (EndoFit OVA; Invi- Differentiated BMMCs were cultured in 24-well plates at a density of 5 voGen) at each treatment site (10 mgin10ml of saline per spot, four spots 5 3 10 cells per well. We used a CW 1064-nm Nd:YVO4 laser for laser in total) on the FITC-painted sites. We then harvested skin-draining treatment as described above. The spot diameter on the well was adjusted lymph nodes (dLN) 24 and 48 h after the vaccination and laser treat- to be ∼16 mm. The BMMC culture was treated with the CW NIR laser at a ment and then isolated DCs from lymph nodes (LN). We mechanically power of 0.5 W/cm2 for 1–3 min. To detect ROS generation, BMMCs were dissociated LN and incubated them with collagenase D (2.5 mg/ml; incubated in 0.5% FBS/HBSS during the laser treatment, then washed and Roche) and 10 mM EDTA as previously described (24). We labeled incubated with 0.1 mMH2DCFDA (Thermo Fisher Scientific) for 15 min cells isolated from LN for CD11c (N418; eBioscience), I-A/I-E (M5/ immediately after the conclusion of the NIR laser treatment. Negative or 114.15.2; eBioscience), CD11b (M1/70; BioLegend), CD103 (2E7; positive control cells were treated with NAC or H2O2, respectively. ROS- BioLegend), langerin/CD207 (4C7; BioLegend), CD8a (53-6.7; eBio- reacted DCF-dependent fluorescence was measured by flow cytometry. science), and CD86 (GL1; BioLegend). We also stained isolated cells We performed data acquisition on a Fortessa cytometer (BD Bioscience) with LIVE/DEAD Fixable Aqua (Life Technologies) to distinguish live followed by analysis on FlowJo v10 software (FlowJo, LLC). Primary cells. We performed data acquisition on a Fortessa cytometer (BD cultured human epidermal keratinocytes were purchased from Life Tech- Bioscience)followedbyanalysisonFlowJosoftware(FlowJo,LLC). nologies. The keratinocytes were cultured in EpiLife media (Life Tech- We selected classical LN-resident cells (classical DCs [cDCs]) based on nologies) with Human Keratinocyte Growth Supplement (Life Technologies) CD11chi status and I-A/I-E intermediate levels. We selected migratory until they reached 80–90% confluency in 24-well plates. We used a CW DCs (migDCs) based on CD11c intermediate levels and I-A/I-Ehi status 1064-nm Nd:YAG laser (Ventus; Laser Quantum, Cheshire, England) for as distinct from cDCs, which were further gated for langerin+ (Lang+) laser treatment. The laser treatment and ROS detection on the kerati- versus CD11b+. We further gated CD11b+ versus CD103+ subpopula- nocyte culture were performed as described above. We performed data + tions from cells within the Lang migDC gate, as previously described acquisition on a BD Accuri C6 Cytometer (BD Bioscience) followed by Downloaded from (20, 24, 25). analysis on the FlowJo software. Fold increase in DCF-positive pop- ulation compared with no laser-treated control was calculated for each Reconstitution of sash mice experiment. We reconstituted MC-deficient sash mice with bone marrow–derived MCs Quantitative PCR analysis of chemokine expression in mouse (BMMCs) as previously described (26). Briefly, we collected bone marrow skin and cultured murine BMMCs (BM) from wild-type (WT) host mice. We cultured BM-derived cell sus- 2 pensions for 6 wk in the presence of 10 ng/ml IL-3 and 10 ng/ml SCF Skin sections measuring 0.2 mm , including both the epidermis and dermis, http://www.jimmunol.org/ (R&D Systems) until we obtained a pure population of MCs. We con- were excised 6 h after laser illumination. The skin tissue was homogenized firmed the differentiation of MCs by flow cytometry with positive staining using a bead mill homogenizer, and total RNA was extracted using the 6 for c-kit and FcεRI. We injected 4 3 10 BMMCs into the flank skin of RNeasy Kit (Qiagen) and reverse transcribed using the RT2 First Strand Kit 4-wk-old sash mice. We allowed mice to recover for 8 wk prior to the (Qiagen). Differentiated BMMCs were cultured in 24-well plates and experiments. treated with the CW NIR laser at a power of 0.5 W/cm2 for 1 min as described above. Four hours after the laser treatment, total RNA was Immunofluorescence analysis of mouse skin extracted and reverse transcribed as described above. The samples were Mouse ears were shaved and depilated 2 d before the tissue preparation. tested for previously identified selected of importance (Ccl2, Ccl6, Depilated ears were treated with the NIR laser as described above. Six hours Ccl11, Ccl17, Ccl20, Il1a) in addition to inflammatory cytokines (Ifng, Tnf, Il1b) (19) using an RT2 qPCR Primer Assay (Qiagen) and RT2 SYBR after the treatment, mice were heart perfused with 4% paraformaldehyde by guest on September 27, 2021 following i.v. injection with DyLight 649-labeled Tomato Lectin (Vector Green ROX qPCR Mastermix (Qiagen) on the StepOnePlus PCR System Laboratories). Lymphatic vessels were identified using Ab to Lyve-1 (1:200, (Applied Biosystems). Fold changes in the expression over sham-treated controls were normalized against housekeeping genes and calculated fol- clone: ALY7; eBioscience; or 1:200, rabbit polyclonal; Relia Tech), 2DDCT and CD11b+ or Lang+ DCs were identified using Ab to CD11b (1:200, lowing the 2 method. Cycle threshold values that were higher than rabbit polyclonal; Novus Biologicals) or langerin/CD207 (1:200, clo- the cutoff of 40 were not considered as reliable and were removed from ne:929F3.01; IMGENEX). We randomly sampled three to four photo- further analysis. 3 2 graphic imaging stacks (450 450 mm , 40–60 mm thick each) for each Quantitative PCR analysis of chemokine expression in cultured ear using confocal laser-scanning microscopy and determined density and width of perfused blood and lymphatic vessels and spatial relationships human keratinocytes + + between CD11b or Lang DCs and blood or lymphatic vessels using Human primary keratinocytes were cultured and treated with the CW NIR a three-dimensional analysis system on Imaris software (Bitplane). We laser at a power of 0.5 W/cm2 for 1 min as described above. Four hours counterstained cell nuclei by DAPI. + after the laser treatment, total RNA was extracted and reverse transcribed CCL21 cells were identified using Ab to CCL21 (1:200, goat poly- as described above. The samples were tested on the Human Inflammatory clonal; R&D Systems). We determined the total surface area of lymphatic Cytokines and Receptors RT2 Profiler PCR Array (number 330231; + vessels, then quantified the number of CCL21 cells on the lymphatic Qiagen) on the GVP-9600 System (Shimadzu, Kyoto, Japan). Fold vessels on Imaris software. These parameters were determined in three to changes in the expression over no treatment controls were analyzed as four photographic areas from each ear. described above. TUNEL staining of mouse skin Measurement of microvascular permeability Mice were heart perfused with 4% paraformaldehyde at 6 h after the Microvascular permeability in the skin tissue was measured as previously treatment with the CW NIR laser at a power of 5.0 W/cm2 for 1 min. described (29). Briefly, Evans blue was injected from retroorbital plexus TUNEL staining was performed on 5-mm-thick paraffin sections using the 0, 1, 2, and 6 h after a 1-min CW 1064-nm laser treatment at 5.0 W/cm2. In Situ Cell Death Detection Kit, TMR red (Roche Applied Science). Thirty minutes after the injection of Evans blue dye, the laser-treated skin Nuclei were identified by counterstaining using DAPI. Slides were imaged tissue portion was excised and dissolved in formamide to elute Evans blue using an inverted fluorescence microscope (Carl Zeiss Axio A1; Carl Zeiss at 55˚C overnight. Evans blue leak in the tissue was quantitated by spec- North America). Apoptotic cells were identified in red and counted in four trophotometry at 650 nm. photographic areas from each skin sample. Statistical analyses Staining of MCs in mouse skin We ran a one-way ANOVA across treatments (vaccine ID only, CW 1064- Mice were heart perfused with 4% paraformaldehyde. Skin biopsies nm laser, PW 1064-nm laser, Alum, c48/80, i.m., NAC) restricted to the WT were embedded in paraffin blocks. Five-micrometer-thick paraffin genotype with Tukey adjusted post hoc pairwise mean comparison tests. We sections were stained with 0.1% toluidine blue (Millipore Sigma) (27). also ran two-way genotype (WT, Ccr72/2, Batf32/2, sash, sash recon- Slides were imaged using an inverted microscope (Axio A1; Carl stituted) by treatment factorial ANOVAs with different selected genotypes Zeiss). Dermal MCs were identified by their characteristic morphology and treatments depending on the question addressed. A log transformation and staining reactions (28) and counted in four to five photographic of Ab titers was applied for Ab titer analysis. Separate analyses were run areas. for the dependent variables IgG, IgG1, and IgG2c. If the interaction was 4 NEAR-INFRARED LASER MODULATES MAST CELLS IN SKIN not significant, we removed it and tested only the main effects of genotype the ID only group in WT mice was observed (Fig. 1A–D). Interest- and treatment. Significant main and interaction effects were followed up ingly, a level of Ab responses in the CW NIR laser group was higher with Tukey post hoc tests. Because the Ab responses of WT mice injected than that in the ID only group in sash-R mice (Fig. 1A–C). Anti- with or without vehicle for NAC treatment were not statistically significant for all the treatments, we pooled the results within the same treatment influenza IgG1 response in the CW NIR group in sash-R mice was groups. significantly higher than that in sash mice (Fig. 1B, p = 0.0259). In For quantitative PCR results, in addition to Student unpaired two-tailed addition, anti-influenza IgG2c response in the CW NIR group in sash- t test with stepdown bootstrap and false discovery rate corrections (30), we R mice was significantly higher than in the ID only group in WT mice ran a discriminant analysis to see if previously identified selected genes of importance (Ccl2, Ccl6, Ccl11, Ccl17, Ccl20, Ifng, Tnf, Il1a, Il1b) (19) (Fig. 1C, p = 0.0347). These results indicate that the attenuated Ab could predict the effect of the NIR laser treatment. response to the CW NIR laser in sash mice was restored by the MC The data analysis for this paper was conducted using SAS/STAT software reconstitution in the skin tissue and that MCs are critical in aug- for Windows (Version 9.4; SAS Institute, Cary, NC) and Prism 7 (GraphPad mentationofAbresponsewiththeCWNIRlaser. software 2016). We used the SAS Candisc Procedure followed by SAS We next assessed the contribution of MCs to protection augmented Stepdisc Procedure for the discriminant analysis. Data were pooled from at least two independent experiments. by the CW NIR laser. The vaccinated mice were challenged intra- nasally with homologous live influenza virus and monitored for survival time. Consistent with the previous reports (18–20), the CW Results NIR laser treatment conferred better protection compared with the ID MCs mediate the adjuvant effect of the CW NIR laser only group (Fig. 1I, 1J). The survival rate in the ID only group in sash MCs are enriched for the expression of genes encoding a diverse array mice was lower than that in WT mice (p = 0.0022), showing a critical of factors involved in the innate and adaptive immune response (31) role of MCs in host defense (32–36). The survival rate in the CW Downloaded from and play a critical role in host defense against parasitic, bacterial, NIR laser in sash mice was significantly lower than that in WT mice and viral (32–36). We therefore hypothesized that the CW (Fig. 1I, 1J, WT versus sash in the CW NIR laser group: p = 0.0030), NIR laser stimulates connective tissue MCs in skin and initiates an suggesting that the effect of the CW NIR laser treatment on protec- immunostimulatory milieu for skin-resident DCs. To test this hy- tion was attenuated in sash mice. In contrast, the CW NIR laser was pothesis, we compared the immune response to an experimental ID able to confer significantly better protection compared with the ID

influenza vaccination (whole inactivated influenza A/PR/8/34 virus) only group in sash-R mice (Fig. 1I, 1J, the CW laser versus ID only http://www.jimmunol.org/ with or without CW or nanosecond PW NIR laser treatment at group in sash-R mice: p = 0.0406), indicating that the effect of the 5.0 W/cm2 for 1 min (18–20) or a MC activator c48/80 (37) in mice. CW NIR laser on protection was restored by the MC reconstitution. Consistent with the previous reports (18–20), the CW, but not PW, These results suggest that MCs play a critical role in augmentation of NIR laser significantly augmented anti-influenza IgG response protective immunity with the CW NIR laser. (Fig. 1A, the CW laser versus ID only group for IgG: p = 0.0033), MCs are known to respond to physical stimuli, including heat (41, provoked a statistically marginal increase in IgG2c (Fig. 1C, the 42). However, consistent with the previous report (20), the PW NIR CW laser versus ID only group for IgG2c: p = 0.0664), induced laser showed marginal immunological effects (Fig. 1A–H), although similar anti-influenza IgG2c:IgG1 ratios (Fig. 1D) without in- it generated the same degree of heat within the skin as did the CW creasing IgE responses (Fig. 1E) and IFN-g and IL-4 secretion NIR laser (19). These observations collectively suggest that heat by guest on September 27, 2021 levels from ex vivo stimulated splenocytes (Fig. 1F, 1G), and in- plays a minimal role in the effect of the CW NIR laser. creased the HAI geometric mean titers (the CW laser versus ID Together, our results support the view that the adjuvant effect of only, 18.81, 95% CI: 9.024–39.19 versus 6.936, 95% CI: 3.520– the CW NIR laser in the context of ID vaccination functionally 13.67) (Fig. 1H) compared with the nonadjuvanted ID only group. depends on MCs in skin. Interestingly, the application of c48/80 significantly augmented anti-influenza IgG response (Fig. 1A, the c48/80 versus ID only MCs mediate augmentation of the migration of skin-resident group for IgG: p = 0.0201) (which is consistent with the published DCs induced by the CW NIR laser literature reporting the adjuvant effect of activated MCs) (38, 39) to MCs are known to significantly impact the adaptive immune re- a similar magnitude as did the CW NIR laser, compared with the ID sponse by regulating trafficking of DCs to dLN (33, 43). We have only group. The geometric mean titers for the c48/80 group were previously shown that exposures of skin to the CW NIR laser en- consistently higher (13.13, 95% CI: 3.824– 45.07) than that of the hance migration of skin-resident migDCs, namely the Lang+ and ID only group, although this difference was not statistically sig- Lang2CD11b2 migDC subsets, although they have no appreciable nificant (Fig. 1H). The c48/80 induced similar anti-influenza IgG2c: effect on cDCs within LN or plasmacytoid DCs, resulting in aug- IgG1 ratios (Fig. 1D) without increasing IgE responses (Fig. 1E) mentation of the adaptive immunity (20). To investigate the role of and IFN-g and IL-4 secretion levels (Fig. 1F, 1G) compared with MCs in the augmentation of migDC migration, we evaluated the the ID only and laser-treated groups, in contrast to alum adjuvant migration response in sash and MC-reconstituted sash-R mice using inducing a profoundly skewed TH2responsewithhighIgEre- the FITC paint technique (23). We applied FITC solution to the sponses (Fig. 1E) and low IFN-g levels (Fig. 1F) and IgG2c:IgG1 back skin of mice prior to the application of the 1-min CW NIR ratios (Fig. 1D) (40). These results show that activation of MCs in laser, followed by ID injection of a model vaccine, OVA. We then skin augments the immune response and produces a mixed TH1and measured the migration of migDC subsets into the skin-dLN by TH2 response. flow cytometry as previously described (20). To further assess the contribution of MCs to the effect of the NIR Consistent with the previous report (20), the CW NIR laser laser, the response in MC-deficient KitW-sh/W-sh (sash) mice and increased the number of FITC+ migDC populations (Fig. 2B, 2C, sash mice selectively reconstituted for MCs in skin (sash-R mice) p , 0.0001, 24 and 48 h) and Lang+ (48 h, p = 0.0341) and by ID injection with genetically compatible BMMCs from WT Lang2CD11b2 (p , 0.0001, 24 and 48 h) migDC subsets (Fig. 2D, mice (26) was also evaluated. In this model, MCs were recon- 2E), as compared with the ID only group. In contrast, the CW stituted locally in the injection sites (Supplemental Fig. 1A–D). NIR laser did not induce any increase in the migDC migration in The effect of the CW NIR laser on anti-influenza Ab responses sash mice, although a comparable level of the migration response was attenuated in sash mice, although a comparable level of Ab to that in the ID vaccine only group of WT mice was observed responses as well as a similar value of IgG2c:IgG1 ratios to those in (Fig. 2B–E). The number of FITC+ migDC and Lang2CD11b2 The Journal of Immunology 5 Downloaded from http://www.jimmunol.org/ by guest on September 27, 2021

FIGURE 1. A critical role of MCs in the effect of the NIR laser on anti-influenza immune responses. The effect of an NIR laser on anti-influenza immune responses was evaluated in WT and MC-deficient sash mice. Reconstitution of sash mice was performed by ID injection of BMMCs 8 wk before vac- cination. WT, sash, and reconstituted mice were vaccinated with 1 mg of inactivated influenza virus (A/PR/8/34) with or without the NIR laser exposure or c48/80 and challenged intranasally with live homologous virus 28 d after vaccination. Immune correlates were analyzed at day 32. Plates for ELISA were coated with the inactivated influenza virus. Serum anti-influenza specific (A) IgG, (B) IgG1, (C) IgG2c, (D) IgG2c/IgG1 ratio, and (E) IgE. (F and G) The effect of the NIR laser on systemic anti-influenza specific T cell responses. The cytokine responses were measured 4 d after challenge by restimulating cultured splenocytes with the inactivated influenza vaccine Ag. Levels of (F) IFN-g and (G) IL-4 in supernatants are (Figure legend continues) 6 NEAR-INFRARED LASER MODULATES MAST CELLS IN SKIN migDC populations of the CW NIR laser group in sash mice was laser, we pretreated mice by s.c. injections of NAC to suppress ROS significantly smaller than that in WT mice (Fig. 2A–E, migDC: generation in response to the CW NIR laser in skin (22) prior to 24 h, p = 0.0031; 48 h, p , 0.0001; Lang2CD11b2 migDC: 24 h, measurements of DC migration, using the FITC paint technique. p = 0.0492; 48 h, p , 0.0001). Interestingly, the CW NIR laser The NAC treatment significantly attenuated the effect of the increased the number of FITC+ migDC populations in sash-R mice CW NIR laser on the migration response of the migDC population compared with that in the ID vaccine only group of WT mice (Fig. 4A) and migDC subsets (Fig. 4B), although a comparable level (Fig. 2A–C, p , 0.0001, 24 h) or the CW NIR laser group of sash of the migration responses in the ID only groups of NAC-treated mice mice (Fig. 2A–C, p = 0.0002, 24 h), indicating restoration of the to that of non–NAC-treated control mice was observed. The mi- migration response of migDCs to the CW NIR laser by the MC gration response of the CW NIR laser-treated group in NAC-treated reconstitution. Consistently, sash-R mice showed the significantly mice was significantly smaller than that in non–NAC-treated mice higher number of FITC+ Lang2CD11b2 migDC population in the (Fig. 4A, migDC: p = 0.0007; Lang2CD11b2: p , 0.0001). CW NIR group compared with those in the ID vaccine only group These results suggest that ROS generation in response to the CW of WT mice (Fig. 2D, 2E, p , 0.0001, 24 h) or the CW NIR laser NIR laser enhances migration of migDCs in skin. group of sash mice (Fig. 2D, 2E, p = 0.0002, 24 h). ROS generation and migDC migration mediate the adjuvant These results indicate that MCs mediate the enhanced migration effect of the CW NIR laser of the migDC subsets induced by the CW NIR laser. Our data suggest that ROS generation followed by migration of the The CW NIR laser induces ROS generation in cultured MCs migDCs plays a pivotal role in the effect of the CW NIR laser. To and keratinocytes probe the contribution of ROS generation upon the CW NIR laser Downloaded from Exposure of skin to NIR light has been reported to induce gen- treatment to the formation of the adaptive immune response, we eration of ROS in the skin tissue (44–47). In turn, ROS have been next treated mice with NAC prior to vaccination and examined the shown to modulate the function of MCs (48). We therefore in- alteration of the immune response to the ID influenza vaccination. vestigated if the CW NIR laser induces ROS generation in MCs, The NAC treatment slightly increased a level of IgG responses as using an established culture model of BMMCs (49) and a ROS- well as IgG2c:IgG1 ratios across the experimental groups, but none

sensitive fluorescence probe H2DCFDA (50). In addition, in light of these increases were significant compared with that in non– http://www.jimmunol.org/ of the fact that ROS can travel across cell types in the tissue (51), NAC-treated control mice (Fig. 5A–D). A level of Ab titers in the we evaluated ROS generation in primary cultured keratinocytes. CW NIR laser group was not higher than that in the ID only group After confirming the maturity and purity of BM-derived cells by flow in NAC-treated mice, whereas an increased level of the Ab re- cytometry (Supplemental Fig. 1A, 1B), cultured BMMCs were treated sponse was observed in the CW NIR laser group compared with with the CW NIR laser for 1–3 min in vitro. ROS-reacted DCF- that in the ID only group in non–NAC-treated control mice dependent fluorescence was measured by flow cytometry. Hydrogen (Fig. 5A–C), indicating that ROS generation is involved in the peroxide–treated cells served as a positive control. DCF fluorescence– adjuvant effect of the CW NIR laser. positive population was identified in the non–laser-treated control cells Our data earlier confirmed the causal link between ROS gener-

(Fig. 3A, 3B), indicating that this system could detect constantly ation and enhanced migration of the migDC subsets in skin in re- by guest on September 27, 2021 generated ROS through aerobic metabolism (52). An addition of an sponse to the CW NIR laser. In addition, we have previously ROS scavenger, NAC, significantly reduced DCF+ cells, confirming demonstrated that the Lang+ DC subset mediates migration of the the constant generation from the control cells (Fig. 3A, 3B). The CW Lang2CD11b2 DC subset in response to the CW NIR laser, using NIR laser treatment significantly increased the ROS-related DCF+ the Lang–diphtheria toxin receptor mouse model (20). However, population in BMMCs (Fig. 3A, 3B, 2.0 folds for 1 min: p = 0.0278; this model cannot specify the distinct roles of epidermal Lang- 2.1 folds for 3 min: p = 0.0157) compared with the non–laser-treated erhans cells and Lang+ dermal DCs in the effect of the CW control group. These data demonstrate that the CW NIR laser induces NIR laser. To further evaluate the contribution of the migration of ROS generation in MCs. The CW NIR laser treatment also increased the migDC subsets via ROS generation to the adjuvant effect, the ROS-related DCF+ population in keratinocytes compared with the we determined the alterations of the immune response to ID vacci- non–laser-treated control group (Fig. 3C, 3D, p = 0.0420). This result nationinBatf32/2 mice. Batf32/2 mice lack the cross-presenting indicates that keratinocytes could serve as a source of ROS for MCs in CD103+ DC subset in skin, which is included in Lang+ dermal response to the CW NIR laser in skin. DCs (53, 54), in mice with the C57BL/6 background (55). Batf32/2 mice have been, therefore, used to test the contribution of the ROS generation with the CW NIR laser mediates the CD103+ migDC subset in several models of skin (56, augmentation of migDC migration 57). Ccr7 loss in Ccr72/2 mice inhibits trafficking of migDC To assess the functional link between the ROS generation and populations from the peripheral tissue (58, 59). The immune re- augmentation of the migDC migration induced by the CW NIR sponse in Ccr72/2 mice was also examined as a control in these

shown. (H) HAI titers 4 d after challenge are shown. (A–D) n = 44, 48, 38, 33, 22, 12, 23, 26, 10, 8, and 10; (E) n = 40, 29, 25, 13, 7, 7, 15, 17, 2, 8, and 10; and (F and G) n = 35, 39, 30, 24, 9, 4, 10, 10, 7, 8, and 10 for no vaccine, vaccine ID only, vaccine ID + CW 1064-nm laser, vaccine ID + PW 1064-nm laser, vaccine + Alum ID, vaccine + c48/80 ID, vaccine ID only in sash, vaccine ID + CW 1064-nm laser in sash, vaccine ID + PW 1064-nm laser in sash, vaccine ID only in sash reconstituted (sash-R), and vaccine ID + CW 1064-nm laser in sash-R groups, respectively. (H) n = 17, 35, 29, 22, 12, and 12 for no vaccine, vaccine ID only, vaccine ID + CW 1064-nm laser, vaccine ID + PW 1064-nm laser, vaccine + Alum ID, and vaccine + c48/80 ID group, respectively. Results were pooled from (A–E)10,(F and G)8,and(H) 5 independent experiments. The p values for the tests of treatments within the WT genotype were based on a one-way ANOVA between subjects of the WT data with Tukey post hoc tests, whereas those for other comparisons are based on a two-way treatment by genotype ANOVAwith Tukey tests. (I) Kaplan–Meier survival plots for 15 d following lethal influenza challenge; Gehan–Breslow–Wilcoxon test. (J) The effect of the NIR laser on body weight of vaccinated mice following viral challenge. Body weights were monitored daily for 15 d. Mean body weight 6 SEM of each experimental group was determined at each time point. n =10, 12, 13, 5, 10, 10, 4, 6, and 6 for no vaccine, vaccine ID only, vaccine ID + CW 1064-nm laser, no vaccine in sash, vaccine ID only in sash, vaccine ID + CW 1064-nm laser in sash, no vaccine in sash-R, vaccine ID only in sash-R, and vaccine ID + CW 1064- nm laser in sash-R groups, respectively. (I and J) Results were pooled from four independent experiments. The Journal of Immunology 7 Downloaded from http://www.jimmunol.org/ by guest on September 27, 2021

FIGURE 2. Effect of the NIR laser on DC migration in skin and the effect of the NIR laser on migration responses of migDC subsets in skin. Mice were painted with 0.5% FITC solution on the flank skin 4 h before vaccination with 40 mg of OVA, with or without the NIR laser treatment. Skin-dLN were analyzed by flow cytometry. (A) Gating schematic to identify DC subsets within skin-dLN. Cell counts of (B and C) DC subsets (D and E) and migDC subsets in the skin-dLN 24 and 48 h after OVAvaccination are shown. (A–D) n = 13, 10, 4, 6, 6, 5, 4, and 4 at 24 h and n = 20, 11, 4, 6, 6, 4, 3, and 5 at 48 h for OVA only in WT, CW 1064-nm laser in WT, PW 1064-nm laser in WT, OVA only in sash, CW 1064-nm laser in sash, PW 1064-nm laser in sash, OVA only in sash-R, and CW 1064-nm laser in sash-R groups, respectively. Results were pooled from six independent experiments and analyzed using two-way ANOVA followed by Tukey honestly significant difference tests. 8 NEAR-INFRARED LASER MODULATES MAST CELLS IN SKIN

FIGURE 3. Effect of the NIR laser on cultured MCs and keratinocytes. (A and B) The effect of the NIR laser on MCs. (A) Release of ROS by the NIR laser treatment was assessed using a ROS-sensitive fluores- cence probe H2DCFDA in BMMCs in vitro. The dif- ferentiated BMMC culture was treated with the NIR laser at a power of 0.5 W/cm2 for 1–3 min. ROS- activated DCF fluorescence was measured by flow cytometry. (B) Fold increase in DCF+ population compared with non–laser-treated control was calcu- lated for each experiment. One-way ANOVA followed by the Tukey honestly significant difference (HSD) tests. n = 6, 8, 12, 7, 6, 6, and 6 for no H2DCFDA control, no laser control, CW 1064-nm laser for 1 min, CW 1064-nm laser for 3 min, no laser control with

NAC, CW 1064-nm laser for 1 min with NAC, and Downloaded from

H2O2-treated positive control groups, respectively. Results were pooled from three independent experi- ments. (C and D) The effect of the NIR laser on pri- mary cultured keratinocytes. (C) Release of ROS in primary cultured human dermal keratinocytes in re- sponse to the CW NIR laser at a power of 0.5 W/cm2 for 1 min was assessed in vitro as described above. (D) http://www.jimmunol.org/ Fold increase in DCF+ population compared with non– laser-treated control was calculated. One-way ANOVA followed by Tukey HSD tests. n = 3 for each of the following groups: no H2DCFDA control, no laser control, CW 1064-nm laser for 1 min, no laser control with NAC, CW 1064-nm laser for 1 min with NAC, and H2O2-treated positive control groups. Repre- sentative data from two independent experiments

are presented. by guest on September 27, 2021

experiments to confirm the importance of migDCs in the forma- mice was observed (Fig. 5A–C). A slight increase in a level of tion of the adaptive immunity to ID vaccination. In addition, we IgG1 responses (Fig. 5B) with a decrease in IgG2c:IgG1 ratios compared the immune responses between ID and i.m. vaccination (Fig. 5D) across the experimental groups in Batf32/2 mice was in these genetic mouse models to assess the contribution of the observed, which is consistent with the previous reports showing + migDC subsets to the immune response augmented with the that the CD103 DC subset is critical for TH1 immunity (54). treatment of skin with the CW NIR laser. None of these changes were significant compared with that in WT Consistent with the previous report (20), Ccr72/2 mice receiving mice. These results indicate that Batf32/2 mice were able to an ID vaccine showed a significant reduction in anti-influenza IgG mount the comparable response to WT mice to ID vaccination, but and IgG2c (Fig. 5A, 5C, WT versus Ccr72/2 in the ID only group: they were not able to respond to the CW NIR laser. p , 0.0001), with similar IgG2c:IgG1 ratios compared with WT IFN-g influenza-specific splenocyte responses were reduced in mice (Fig. 5D). In addition, IFN-g influenza-specific splenocyte Batf32/2 mice as compared with those in WT mice (Fig. 5E, WT responses in Ccr72/2 mice were significantly reduced, as compared versus Batf32/2 in the vaccine ID only group: p , 0.0001; in the with those in WT mice (Fig. 5E, WT versus Ccr72/2 in the vaccine CW laser group: p , 0.0071), whereas IL-4 responses were not ID only group: p , 0.0001, WT versus Ccr72/2 in the CW NIR significantly changed (Fig. 5F). In contrast, i.m. administration of laser group: p , 0.0001). A level of Ab responses in the CW NIR the same dose of the vaccine mounted a comparable level of anti- laser group in Ccr72/2 mice showed a significant reduction for IgG influenza Ab responses across the IgG subclasses among WT, (Fig. 5A, p , 0.0001) and statistically marginal reduction for Ccr72/2, and Batf32/2 mice (Fig. 5A–C), highlighting a critical IgG2c (Fig. 5C, p = 0.0693) compared with that in WT mice. These role of the dermal CD103+ DC subset in the effect of the CW NIR results support dependence of the response to the ID vaccination laser on the Ab response in the context of the ID vaccination. with or without the CW NIR laser treatment on migDC, as Ccr72/2 These results collectively demonstrate that the dermal CD103+ mice lack CCR7-dependent migration of migDCs (59). DC subset is critical to the adjuvant effect of the CW NIR laser, as Interestingly, a level of Ab titers in the CW NIR laser group was opposed to epidermal Langerhans cells playing a predominant role not higher than that in the ID only group in Batf32/2 mice, whereas in the adjuvant effect of the visible PW laser adjuvant (15). a comparable level of the Ab response of the ID only and CW NIR We next assessed the contribution of the dermal CD103+ DC laser groups in Batf32/2 mice to that of the ID only group in WT subset to protection augmented by the CW NIR laser. Mice were The Journal of Immunology 9

FIGURE 4. The impact of ROS deletion on DC migration enhanced by the NIR laser. The effect of ROS deletion on the migration response of DCs to the NIR laser in skin. Mice were treated with s.c. injection of 100 mg/kg NAC daily for four consecutive days before vaccination and the NIR laser treatment,as appropriate. Mice were painted with 0.5% FITC solution on the flank skin 4 h before vaccination with 40 mg of OVA with or without the NIR laser treatment. Cell counts of (A) DC subsets and (B) migDC subsets in the skin-dLN 48 h after OVA vaccination are shown. Results were pooled from four Downloaded from independent experiments. One-way ANOVA followed by Tukey honestly significant difference tests. n = 20, 11, 6, and 7 for OVA only in WT, CW 1064-nm laser in WT, OVA only in NAC-treated groups, and CW 1064-nm laser in NAC-treated groups, respectively. Data of the control group from Fig. 2 are shown for comparison. challenged intranasally with homologous live influenza virus and microvascular permeability was observed in sash mice (Fig. 6C), monitored for survival time as described above. The survival rate demonstrating that the increase in microvascular permeability in- http://www.jimmunol.org/ of unvaccinated Batf32/2 mice was lower than that of WT mice duced by the CW NIR laser was mediated by MCs. (Fig. 5G, p = 0.0134), showing a critical role of the dermal CD103+ The endothelial dysfunction caused by damage may induce an DC subset in host defense (55). Consistent with the Ab response increase of microvascular permeability (60). The laser param- study (Fig. 5A–C), the survival rate in the CW NIR laser in Batf32/2 eter for the CW NIR laser treatment used in this study has mice was significantly lower than that in WT mice (Fig. 5G, WT been demonstrated to be non–tissue damaging in our previous versus Batf32/2 in the CW NIR laser group: p = 0.0003), with studies (18–20). Consistently, the laser treatment did not greater weight loss upon viral challenge (Fig. 5H), suggesting that significantly increase the number of TUNEL+ cells in skin the beneficial effect of the CW NIR laser on protection against (Supplemental Fig. 2). lethal influenza virus challenge observed in WT mice was atten- These results confirm that the CW NIR laser activates MCs and by guest on September 27, 2021 uated in Batf32/2 mice. These results confirm an important role of induces functional changes in microcirculation in skin. the CD103+ migDC subset in the beneficial effect of the CW NIR The CW NIR laser induces morphological changes in laser on protection. lymphatic vessels and migration of migDCs to the lymphatics In conclusion, these results support the view that ROS generation in response to the CW NIR laser modulates the MC function and Microcirculation including blood vessels and the lymphatic net- enhances the migration of the CD103+ migDC subset in skin, work plays an indispensable role in recruitment of immune cells to ultimately augmenting the immune response to ID vaccination. inflammatory sites, transportation of pathogenic Ags into dLN, and migration of Ag-presenting and effector cells to secondary lym- The CW NIR laser induces functional changes in phoid organs (61). We next determined the response of the vas- microcirculation in skin cular and lymphatic networks in skin to the CW NIR laser without MCs have been shown to induce functional changes in microcir- vaccination. Mice received exposures of the CW NIR laser on the culation via their secretory responses, regulating vascular flow and ear skin tissue at 5.0 W/cm2 for 1 min. Six hours later, we per- permeability and recruitment of inflammatory cells (31). We formed immunofluorescence of the whole-mount ear tissue and therefore hypothesized that the CW NIR laser establishes an stained the lymphatic network and skin-resident DCs for analysis immunostimulatory milieu to augment protective immune re- by confocal imaging. sponses in microcirculation in skin via MC activation. To test The CW NIR laser treatment increased the density of the this hypothesis, we first measured microvascular permeability lymphatic network (Fig. 7A–C, the non–laser-treated versus laser- in the skin tissue treated with the CW NIR laser in WT and treated group for lymphatic density: p = 0.0008), suggesting MC-deficient sash mice. increased lymphatic flow in the laser-treated tissue, which is The CW, but not PW, NIR laser treatment increased microvascular typically induced by an increase in microvascular permeability permeability immediately after the treatment (Fig. 6A, 6B, the non– (62). Interestingly, this functional change was accompanied by a laser-treated versus laser-treated group at 0 h: p =0.0305).The concomitant decrease in the distance between lymphatic vessels increase in microvascular permeability was no longer observed 1 h and Lang+, but not CD11b+, DCs (Fig. 7D, the non–laser-treated after the treatment, suggesting that the effect of the NIR laser on versus laser-treated group for Lang+ DC: p = 0.0302), indicating microvascular permeability resolved within 1 h. The PW NIR laser that the functional changes in the lymphatics provided migration did not induce applicable increase, which confirms that heat gen- cues for skin-resident DCs. The CW NIR laser treatment without eration in skin plays a minimum role in the functional changes vaccination induced no change in the number of migDCs in the induced by the CW NIR laser. To investigate the role of MCs in treated skin (Fig. 7E) nor in skin-dLN, which was examined using the transient increase in microvascular permeability, we further FITC painting technique (Fig. 7F, 7G), suggesting that Ag chal- assessed the effect of the CW NIR laser in sash mice. No increase in lenge is necessary to induce the migDC migration into LN (63). 10 NEAR-INFRARED LASER MODULATES MAST CELLS IN SKIN Downloaded from http://www.jimmunol.org/ by guest on September 27, 2021

FIGURE 5. The impact of ROS deletion or selective ablation of DC subpopulations on the immune responses augmented by the NIR laser. The effect of the NIR laser on the immune response to the ID influenza vaccination was evaluated in NAC-treated mice or Ccr72/2 and Batf32/2 mice. For antioxidant treatment, mice were treated with NAC for four consecutive days before vaccination and the NIR laser treatment as described above. Serum anti-influenza specific (A) IgG, (B) IgG1, (C) IgG2c, and (D) IgG2c/IgG1 ratio 4 d after challenge are shown. The p values for the tests of treatments within the WT genotype were based on a one-way ANOVA between subjects of the WT data with Tukey post hoc tests, whereas those for other comparisons are based on a two-way treatment by genotype ANOVAwith Tukey tests. (E and F) Systemic anti-influenza specific T cell responses measured in restimulated splenocytes upon influenza vaccination. Levels of (E) IFN-g and (F) IL-4 in supernatants are shown. (A–D) n = 44, 48, 38, 33, 25, 11, 12, 11, 14, 14, 3, 8, 9, 9, 3, and 6 and (E and F) n = 35, 39, 30, 24, 16, 11, 12, 11, 14, 14, 3, 8, 9, 9, 3, and 6 for no vaccine, vaccine ID only, vaccine ID + CW 1064-nm laser, vaccine ID + PW 1064-nm laser, vaccine i.m., vaccine ID only in NAC-treated groups, vaccine ID + CW 1064-nm laser in NAC-treated (Figure legend continues) The Journal of Immunology 11

A FIGURE 6. The effect of the NIR laser on microvascular permeability in skin. ( ) Representative images of the mouse back skin 30 min after i.v. Downloaded from injection of Evans blue. (B) Quantification of tissue concentration of Evans blue 0, 1, 2 and 6 h after the 1-min CW or PW NIR 1064-nm laser treatment is shown. n = 8, 12, and 6 for no laser, CW 1064-nm laser, and PW 1064-nm laser groups, respectively, at 0 h, and n = 3, 3–4, and 3–4 for no laser, CW 1064- nm laser, and PW 1064-nm laser groups, respectively, at 1, 2, and 6 h. One-way ANOVA followed by Tukey honestly significant difference (HSD) tests. (C), The effect of the NIR laser on microvascular permeability in skin in WT and sash mice. Quantification of tissue concentration of Evans blue 0 h after the 1-min CW NIR 1064-nm laser treatment in WT and sash mice. n = 8, 12, 3, 5, and 5 for no laser in WT, CW 1064-nm laser in WT, no laser in sash, CW 1064-nm laser in sash, and c48/80 in WT mice groups, respectively. Two-way ANOVA followed by Tukey HSD tests. WT data of the no laser group at 0 h from (B) are shown for comparison. (A–C) Results were pooled from three independent experiments. http://www.jimmunol.org/

Together, these results demonstrate that the CW NIR laser in- mice (Fig. 8A, the non–laser-treated versus laser-treated group in duces functional changes in the lymphatic network and provides WT: p = 0.0437; Supplemental Fig. 3). Consistent with the previous migration cues for skin-resident DCs toward the lymphatics in skin. report (19), no upregulation of inflammatory cytokines including Tnf and Il1b was observed. This change appeared to be transient in The CW NIR laser induces a selective set of chemokine nature, as the upregulation of the genes of interest was not observed upregulation in skin at 2 or 24 h (Fig. 8B). Importantly, no significant increase in the

The CW NIR laser appears to induce a transient increase in mi- selective chemokine expression was observed in sash mice with the by guest on September 27, 2021 crovascular permeability and lymphatic flow, which is consistent CW NIR laser treatment (Fig. 8A), suggesting that these changes in with the tissue changes upon MC activation (64). Chemical agents, chemokine expression were dependent on MCs. which are designed to stimulate MC function, generally induce Our data suggest that ROS generation plays a critical role in the prolonged and exacerbated inflammatory responses (65). In con- adjuvant effect of the CW NIR laser. To explore the possible causal trast, exposures of skin to a low-power CW NIR laser have been link between ROS generation upon the CW NIR laser treatment shown to be free from tissue damage or inflammation, evidenced and chemokine expression, we next treated mice with NAC and by no polymorphonuclear infiltration (18–20) with upregulation of examined the alteration of the chemokine response. The NAC a selective set of chemokines (19). We therefore hypothesized that treatment attenuated the effect of the CW NIR laser on expression the CW NIR laser induces “differential” or “selective” secretion of of the selective set of chemokines (Fig. 8A), indicating that ROS mediators without degranulation of MCs (66, 67). To test this generation mediates the chemokine expression in skin. hypothesis, we assessed expression of a selective set of chemo- To further uncover pathways involved in the ROS generation and kines, which was established previously (19), as well as inflam- chemokine responses, we treated cultured BMMCs or keratino- matory cytokines in WT and MC-deficient sash mice. cytes with the CW NIR laser and measured expression of the The CW, but not PW, NIR laser treatment induced temporary selective set of chemokines in BMMCs or evaluated an expression expression of the selective chemokines at 6 h. Student t test with array of inflammatory cytokines and receptors in keratinocytes stepdown bootstrap and false discovery rate corrections showed in vitro. Except for a slight increase noted in Ccl11, Ccl17, and that none of the nine selected chemokines and cytokines of im- Ccl20 in BMMCs, no significant change was observed in the other portance predicted the effect of the NIR laser treatment signifi- chemokines in BMMCs or 84 genes examined on the array in cantly. However, a multivariate discriminant analysis that tests the keratinocytes (Supplemental Fig. 4). These results suggest that linear combination of the nine genes, which optimally discrimi- the skin microenvironment or collaboration among skin cells is nates the treatments, found a significant treatment effect in WT essential to orchestrate the response to the CW NIR laser.

groups, vaccine ID + PW 1064-nm laser in NAC-treated groups, vaccine ID only in Ccr72/2 groups, vaccine ID + CW 1064-nm laser in Ccr72/2 groups, vaccine ID + PW 1064-nm laser in Ccr72/2 groups, vaccine i.m. in Ccr72/2 groups, vaccine ID only in Batf32/2 groups, vaccine ID + CW 1064-nm laser in Batf32/2 groups, vaccine ID + PW 1064-nm laser in Batf32/2 groups, and vaccine i.m. in batf32/ groups, respectively. (A–F) WT data from Fig. 1 are shown for comparison. Results were pooled from six independent experiments. (G and H) The effect of the NIR laser on protection. (G) Kaplan–Meier survival plots. (H) The effect of the NIR laser on body weight following viral challenge. Body weights were monitored daily for 15 d. Mean body weight 6 SEM of each experimental group was determined at each time point. (G and H) n = 10, 12, 13, 4, 5, and 8 for no vaccine in WT, vaccine ID only in WT, vaccine ID + CW 1064-nm laser in WT, no vaccine in Batf32/2, vaccine ID only in Batf32/2, and vaccine ID + CW 1064-nm laser in Batf32/2 groups, respectively. Gehan–Breslow–Wilcoxon test. Results were pooled from three independent experiments. WT data from Fig. 1 are shown for comparison. 12 NEAR-INFRARED LASER MODULATES MAST CELLS IN SKIN Downloaded from http://www.jimmunol.org/ by guest on September 27, 2021

FIGURE 7. The effect of the NIR laser on microcirculation and DC migration without Ag challenge in skin. (A–E) Morphometry of the vascular and lymphatic networks in skin with or without the NIR laser treatment. Morphological changes of the vascular and lymphatic networks in skin were measured in stacks of the three-dimensional confocal images using Imaris software. The depilated mouse ears were treated with the NIR laser, and the whole-mount ear tissue was immunostained and imaged for blood vessels, lymphatics, and DC subsets 6 h after the treatment. (A) Representative confocal images of mouse ear stained for blood vessels, lymphatics, and DCs 6 h following the CW NIR laser treatment. Scale bar, 40 mm. (B and C) Morphometry of blood and lymphatic vessels and DCs in skin using Imaris software. Two-way ANOVA followed by Tukey honestly significant difference (HSD) tests. (A–C) n =17 and 17 for no laser control and CW NIR laser-treated groups, respectively. (D) Distances between vasculature and DCs (Figure legend continues) The Journal of Immunology 13

Together, these results further support the view that the CW NIR expensive, posing financial and engineering challenges including laser induces transient and selective activation of MCs, possibly in high cost per unit and needs of maintenance by laser specialists. In concert with other types of skin cells, via ROS generation. contrast, the low-wattage CW NIR lasers are a mature and simple technology, and a small device emitting a CW NIR laser at the The CW NIR laser induces CCL21 upregulation in required powers used in this application (5 W/cm2 and below) the lymphatics proved to be economically produced (18), making this technology Migration of DCs is a critical step for the initiation of protective feasible and suitable for mass-vaccination programs and other immune responses (61). In response to inflammatory stimuli, the clinically applicable approaches. lymphatic network is known to increase expression of CCL21 on In this study, we have shown that ROS generation in MCs is lymphatic vessels, triggering migration of skin-resident DCs (68, critical for the effect of the CW NIR laser. There is ample evidence 69). Because our earlier observations showed that the CW NIR laser showing that NIR light induces generation of ROS in exposed tissue induced migrational changes of migDCs, we further hypothesized or cells. Low-power (1 mW–5 W/cm2) NIR light between 600 and that the immunostimulatory milieu established by the selective ac- 1000 nm has been established to show diverse beneficial effects in tivation of MCs with the CW NIR laser induces CCL21 response in exposed tissue, including pain relief, facilitation of tissue regen- the lymphatics. To test this hypothesis, we performed immuno- eration, and reduction of inflammation. These biological effects fluorescence of CCL21 and the lymphatic network in the mouse ear of NIR light are broadly defined as photobiomodulation (44, tissue exposed to the CW NIR laser for analysis by confocal 71–76). Although the precise mechanisms of action for photo- imaging. To evaluate the contribution of MCs, we also assessed biomodulation are still under debate (44, 75–77), it is clear that the

CCL21 response in MC-deficient sash mice. effect of NIR light is partially, if not entirely, mediated by ROS Downloaded from CCL21 expression on Lyve-1+ lymphatics was distributed among (44, 73–76). In turn, there is ample evidence showing that NIR small patch structures in the control and exposed skin tissue (Fig. 9A), light induces generation of ROS in exposed tissue or cells. Ex- as previously reported (70). Image analysis revealed an increase in the posures to NIR light between 760 and 1400 nm have been shown density of CCL21+ cells in the exposed skin tissue compared with the to increase ROS generation in tumor cells (78–80), keratinocytes nonexposed control tissue (Fig. 9B, p = 0.0314), which appeared to (79), fibroblasts (81), and the skin tissue (46, 47, 82). In contrast to

mediate the chemotaxis of migDCs to lymphatic vessels. Importantly, photobiomodulation, ROS generation upon exposure to NIR light http://www.jimmunol.org/ the effect of the CW NIR laser on the CCL21 response was not ob- has been linked to deleterious effects, including premature skin served in sash mice (Fig. 9A, 9B). In the current study, we have shown aging (83, 84) and the cytotoxic effect (85, 86). Beneficial or that ROS are generated in MCs and keratinocytes in response to the deleterious, ROS have been postulated to be generated by mito- CW NIR laser (Fig. 3A–D). Thus, it is possible that ROS from various chondrial cytochrome c oxidase in the respiratory electron trans- sources in skin directly stimulate lymphatic endothelial cells and induce port chain in response to NIR light exposure (87, 88), ultimately CCL21 expression. However, these data indicate that MCs are required activating the NF-kB signaling pathway and innate signaling (44, for the upregulation of CCL21 in response to the CW NIR laser and 73–75). In contrast, the numerous reliable data have reported that that mediators derived from activated MCs via ROS generation are ROS play a significant role in MC-dependent inflammatory critically involved in the CCL21 upregulation. processes and activation of intracellular signaling for production by guest on September 27, 2021 In summary, these results show that the CW NIR laser induces of proinflammatory cytokines that regulate the innate immune CCL21 expression in the lymphatics in skin via MC activation and response (48, 89). In accordance, we found that the CW NIR laser confirm our earlier observations that the CW NIR laser provides induced ROS generation in MCs (Fig. 3) and that the effects of the migratory cues for migDCs. CW NIR laser on the migration of migDCs and the adaptive im- mune response are functionally dependent on ROS generation Discussion (Figs. 4, 5). In addition, in the absence of MCs, we observed no In this study, we have shown for the first time, to our knowledge, upregulation of the selective set of chemokines (Fig. 8) nor sub- that a CW NIR laser with a distinct power and dose has a unique sequent augmentation of the immune response by the CW NIR capability to reproducibly activate selective innate programs, including laser (Fig. 1). These results clearly indicate that MCs locate at an MC function and migDC migration, via ROS generation without intersection between the adaptive immune response and ROS gen- inducing prolonged and potentially harmful inflammatory responses eration in response to the CW NIR laser. However, the precise mode in skin, thus augmenting the adaptive immune response (Fig. 9C). of action for the MC activation via ROS generation upon the CW These mechanistic findings would significantly contribute to opti- NIR laser treatment has yet to be defined. For example, it is possible mization of wavelength, power, duration of the treatment, the tim- that ROS generated in adjacent cells reach MCs in skin, modulating ing and location of delivery, and the size of the treatment spot on the MC function indirectly, as ROS are small enough to travel across skin of a CW NIR laser adjuvant in the context of ID vaccination cell types in tissues (51). Further studies are warranted to determine for better safety and efficacy of this technology. the precise molecular identify of ROS, subcellular localization of Although other medical PW lasers have been similarly reported ROS generation, spatial and temporal dynamics of ROS distribution to show the adjuvant effect to ID vaccines (10, 11), use of a CW in tissues, and the mode of activation of innate signaling pathways NIR laser has clear advantages over the other classes of lasers. by ROS in response to the CW NIR laser treatment to selectively Current clinical-grade PW laser systems in these reports are typ- activate MCs and induce the specific immunomodulatory effect ically large in size, complicated in machinery to generate PW, and without deleterious inflammation.

and (E) density of Lang+ and CD11b+ DCs in skin. n = 5–8 and 5–8 for no laser control and CW NIR laser-treated groups, respectively. Two-way ANOVA followed by Tukey HSD tests. (A–E) Results were pooled from five independent experiments. (F and G) The effect of the NIR laser on the migration response of migDC subsets without Ag challenge in skin. Mice were painted with a 0.5% FITC solution on the flank skin 4 h before the NIR laser treatment without any Ag challenge. Cell counts of (F) DC subsets and (G) migDC subsets in the skin-dLN 24 and 48 h after FITC painting are shown. (F and G) n = 13–14, 7, and 7 for no laser, CW 1064-nm laser, and PW 1064-nm laser groups, respectively. Two-way ANOVA followed by Tukey HSD tests. Results were pooled from four independent experiments. 14 NEAR-INFRARED LASER MODULATES MAST CELLS IN SKIN

1064-nm laser at a frequency of 10 kHz did not show any sig- nificant immunological effect, suggesting that this particular parameter is suboptimal in inducing proper ROS generation for the subsequent immunological events. An optimal parameter of PW NIR laser might be sought for in the context of the mo- lecular identify, local sources, and subcellular localization of ROS in response to the NIR laser ultimately leading to efficient activation of innate signaling in the future study. IR irradiation and heat generation in exposed tissues have been linked to biological effects, including premature skin aging via expression of MMP-1 or generation of ROS (94). Accordingly, heat generated by NIR exposures could have a predominant role in MC activation and augmentation of the immune response. In fact, MCs are capable of responding to physical stimuli through transient receptor potential channels, which are functionally coupled to calcium signaling and regulate mediator release in MCs (42). Moreover, transient receptor potential channels of the vanilloid subtype have been shown to respond to IR radiation

(41). However, our study shows that a thermal mechanism is not Downloaded from a major driver of the impact of the NIR laser on the . As the CW and PW 1064-nm lasers in this study were emitted from the same platform with the same wavelength and averagepower,theyprovedtohavethesameheatgeneration profile in skin (19). Despite the same thermal profile, the PW

NIR laser has shown little impact on MC function in skin http://www.jimmunol.org/ (Fig. 6), if any, resulting in the marginal effect on the immune response (Fig. 1) (20). It is still possible that heat is required for the CW NIR laser to have these effects, as one study shows that water-filtered NIR light between 760 and 1400 nm induces ROS generation in a temperature-dependent manner (82). Nonethe- less, our data support the view that heat alone is not sufficient to induce the subsequent immunological events and that specific FIGURE 8. A critical role of MCs and ROS in the chemokine expres- sion in skin in response to the NIR laser. The effect of the NIR laser on the photochemical events related to the CW NIR laser exposure are by guest on September 27, 2021 chemokine expression in skin was evaluated in WT and sash mice. The role required for these effects. of ROS in the chemokine expression in skin in response to the CW NIR Cytokines and chemokines are released by MCs, keratinocytes, laser was also evaluated in NAC-treated WT mice. For antioxidant treat- DCs, melanocytes, tissue-resident macrophages, and inflammatory ment, mice were treated with NAC for four consecutive days before the infiltrates in skin (95). The cutaneous cytokine and chemokine NIR laser treatment as described above. (A) The expression of chemokines network is critically involved in regulation of skin functions as in the mouse back skin was measured 6 h following the CW NIR laser well as inflammatory conditions such as psoriasis or atopic der- treatment in (left) WT, (middle) sash, and (right) NAC-treated WT mice, matitis (95–98). Because of their anatomical proximity, collabo- n using qPCR. = 7, 7, 4, 3, 3, 3, and 3 for no laser control in WT, CW, and rative activities between MCs and keratinocytes via cytokine and PW 1064-nm laser treatment in WT groups; no laser control in sash, CW chemokine signaling have been reported in coordinating cell 1064-nm laser treatment in sash, NAC-treated only, and CW 1064-nm laser NAC-treated groups, respectively. (B) The expression of chemokines in the proliferation and differentiation (96, 99–102) and the immune mouse back skin in (left) 2 and (right) 24 h after the NIR laser treatment in response and inflammation (95, 103). The chemokines and their WT mice. n = 4 and 4 at 2 h, and n = 3 and 3 at 24 h for no NIR and CW receptors upregulated in response to the CW NIR laser have been 1064-nm laser groups, respectively. (A and B) Student t test with stepdown reported to be expressed in both MCs (103) and keratinocytes (95, bootstrap and false discovery rate corrections followed by a multivariate 104) under certain conditions. In the current study, we have shown discriminant analysis of the nine genes was used as appropriate. Error bars that the chemokine response to the CW NIR laser in the mouse show means 6 SEM. skin is dependent on ROS generation and MCs (Fig. 8). Al- though ROS generation was observed in cultured BMMCs and Our data collectively demonstrate that there are distinct dif- keratinocytes in vitro (Fig. 3), small changes were detected in the ferences in the immunological effect between CW and PW NIR subsequent chemokine expression in the individual culture sys- lasers. Although both CW and PW NIR lasers have been reported tems (Supplemental Fig. 4). These results suggest that the che- to induce ROS generation in the biological system (46, 47, 78– mokine response to the CW NIR laser is dependent on the local 82, 90, 91), the exact molecular identify, temporal distribution, skin microenvironment and interactions among MCs, keratino- or local concentration of ROS leading to the immunological cytes, and possibly other types of cells, including fibroblasts in events in response to the NIR lasers has not been characterized. close association with MCs and keratinocytes in skin. Further A PW NIR laser has been shown to be more effective in ROS investigation is warranted to uncover the cross-talk among these generation (90, 91) and photobiomodulation (92, 93) than CW cells in the context of regulation of this complex skin chemokine in some biological systems. Because the pulse frequency, power network in response to the CW NIR laser. level, and exposure time of an NIR laser can dictate a level of Our data show that the CW NIR laser induces a transient increase ROS generation in tissues (90, 91), we might be able to identify in microvascular permeability and lymphatic flow, which is con- a parameter of PW NIR laser for the desired immunological sistent with the tissue changes upon MC activation (64). Contrary to effect. Nonetheless, in the current study, a nanosecond-pulsed chemical agents that are designed to stimulate MC function and The Journal of Immunology 15 Downloaded from http://www.jimmunol.org/ by guest on September 27, 2021

FIGURE 9. The effect of the NIR laser on CCL21 expression on the lymphatics in skin. Quantification of CCL21 expression on the lymphatics in stacks of the three-dimensional confocal images using Imaris software. The depilated mouse ears were treated with the NIR laser, and the whole-mount ear tissue was stained and imaged for the lymphatics and CCL21 6 h after the treatment. (A) Representative confocal images of CCL21 immunofluorescence are shown. Scale bar, 40 mm. (B) Quantification of CCL21+ cells on Lyve-1+ lymphatic vessels is shown. n = 6, 6, 3, and 3 for no laser control in WT (CD11c- EYFP), CW 1064-nm laser treatment in WT (CD11c-EYFP), no laser control in sash, and CW 1064-nm laser in sash groups, respectively. Two-way ANOVA followed by Tukey honestly significant difference tests. Results were pooled from three independent experiments. (C) Schematic diagram of the mechanism of action of the CW NIR laser adjuvant. may cause prolonged inflammatory responses with off-target effects In summary, our results show that a CW NIR laser possesses a (65), the low-power NIR exposures of skin are free from tissue unique ability to selectively modulate innate responses in skin. damage or inflammation (18–20). This feature may be compatible These findings built upon mechanistic knowledge of the biological with differential or selective secretion of mediators without de- effect of the CW NIR laser would not only significantly contribute granulation of MCs (66, 67). Unlike allergic reactions, MCs have to optimization of the NIR laser vaccine adjuvant but also allow been rarely reported to degranulate during inflammatory processes exploration of a novel use of this technology for the treatment of (66, 67). This activation in the nonallergic processes could be immune-related diseases. explained by selective secretion of mediators without degranulation through a process associated with their ultrastructural alterations Acknowledgments (105–108). This hypothesis is further supported by the observations We thank the Live Cell Imaging Facility, Karolinska Institutet, Sweden, the that the CW NIR laser induces a selective set of chemokines Swedish Research Council, the Centre for Innovative Medicine and the Jon- without upregulating inflammatory cytokines, including TNF-a asson Center at the Royal Institute of Technology, Sweden, and the Collabo- and IL-1b, and that the upregulation in the exposed skin tissue rative Research Resources, Keio University School of Medicine, Tokyo, Japan is dependent on MCs (Fig. 8). Reproducible control of micro- for the use of Imaris software, which provided the three-dimensional confocal image analysis service. We also thank Dr. Thomas J. Diefenbach (Ragon In- vascular permeability or lymphatic flow without widespread in- stitute of MGH, MIT, and Harvard) for expert technical assistance with con- flammation by CW NIR laser may offer a desirable approach for focal imaging, the Ragon Institute Imaging Core Facility for three-dimensional other medical purposes, including facilitation of drug delivery or confocal imaging, the MGH Department of Pathology Flow and Image Cytom- the treatment of . An optimal parameter of a CW NIR etry Research Core for flow cytometry, and Madeline Penson and Don Sobell laser for such an application is yet to be found and awaits vali- (all in the Vaccine and Immunotherapy Center at MGH) for excellent technical dation in relevant models in the future. assistance and fruitful discussions. 16 NEAR-INFRARED LASER MODULATES MAST CELLS IN SKIN

Disclosures 26. Grimbaldeston, M. A., C. C. Chen, A. M. Piliponsky, M. Tsai, S. Y. Tam, and S. J. Galli. 2005. Mast cell-deficient W-sash c-kit mutant Kit W-sh/W-sh mice The authors have no financial conflicts of interest. as a model for investigating mast cell biology in vivo. Am. J. Pathol. 167: 835–848. 27. Gaudenzio, N., R. Sibilano, P. Starkl, M. Tsai, S. J. Galli, and L. L. Reber. 2015. Analyzing the functions of mast cells in vivo using ‘mast cell knock-in’ References mice. J. Vis. Exp. 99: e52753. 1. Fehres, C. M., J. J. Garcia-Vallejo, W. W. Unger, and Y. van Kooyk. 2013. Skin- 28. Wershil, B. K., Z. S. Wang, J. R. Gordon, and S. J. Galli. 1991. Recruitment of resident antigen-presenting cells: instruction manual for vaccine development. neutrophils during IgE-dependent cutaneous late phase reactions in the mouse Front. Immunol. 4: 157. is mast cell-dependent. Partial inhibition of the reaction with antiserum against 2. Sticchi, L., M. Alberti, C. Alicino, and P. Crovari. 2010. The intradermal tumor necrosis factor-alpha. J. Clin. Invest. 87: 446–453. vaccination: past experiences and current perspectives. J. Prev. Med. Hyg. 51: 29. Radu, M., and J. Chernoff. 2013. An in vivo assay to test blood vessel per- 7–14. meability. J. Vis. Exp. 73: e50062. 3. Combadiere, B., and C. Liard. 2011. Transcutaneous and intradermal vacci- 30. Westfall, P. H., and S. S. Young. 1993. Resampling-Based Multiple Testing: nation. Hum. Vaccin. 7: 811–827. Examples and Methods for p Value Adjustment. Wiley, New York. 4. Icardi, G., A. Orsi, A. Ceravolo, and F. Ansaldi. 2012. Current evidence on 31. Immunological Genome Project Consortium. 2016. Expression profiling of intradermal influenza vaccines administered by Soluvia™ licensed micro in- constitutive mast cells reveals a unique identity within the immune system. Nat. jection system. Hum. Vaccin. Immunother. 8: 67–75. Immunol. 17: 878–887. 5. Principi, N., L. Senatore, and S. Esposito. 2015. Protection of young children 32. Wong, G. W., L. Zhuo, K. Kimata, B. K. Lam, N. Satoh, and R. L. Stevens. from influenza through universal vaccination. Hum. Vaccin. Immunother. 11: 2014. Ancient origin of mast cells. Biochem. Biophys. Res. Commun. 451: 314– 2350–2358. 318. 6. Young, F., and F. Marra. 2011. A systematic review of intradermal influenza 33. Abraham, S. N., and A. L. St John. 2010. Mast cell-orchestrated immunity to vaccines. Vaccine 29: 8788–8801. pathogens. Nat. Rev. Immunol. 10: 440–452. 7. Bonam, S. R., C. D. Partidos, S. K. M. Halmuthur, and S. Muller. 2017. An 34. Galli, S. J., and M. Tsai. 2008. Mast cells: versatile regulators of inflammation, overview of novel adjuvants designed for improving vaccine efficacy. Trends tissue remodeling, host defense and homeostasis. J. Dermatol. Sci. 49: 7–19. Pharmacol. Sci. 38: 771–793. 35. Urb, M., and D. C. Sheppard. 2012. The role of mast cells in the defence Downloaded from 8. Vitoriano-Souza, J., N. Moreira, A. Teixeira-Carvalho, C. M. Carneiro, against pathogens. PLoS Pathog. 8: e1002619. F. A. Siqueira, P. M. Vieira, R. C. Giunchetti, S. A. Moura, R. T. Fujiwara, 36. Graham, A. C., R. M. Temple, and J. J. Obar. 2015. Mast cells and influenza a M. N. Melo, and A. B. Reis. 2012. Cell recruitment and cytokines in skin mice virus: association with allergic responses and beyond. Front. Immunol. 6: 238. sensitized with the vaccine adjuvants: saponin, incomplete Freund’s adjuvant, 37. Paton, W. D. 1951. Compound 48/80: a potent histamine liberator. Br. J. and monophosphoryl lipid A. PLoS One 7: e40745. Pharmacol. Chemother. 6: 499–508. 9. Chen, X., M. Pravetoni, B. Bhayana, P. R. Pentel, and M. X. Wu. 2012. High 38. McLachlan, J. B., C. P. Shelburne, J. P. Hart, S. V. Pizzo, R. Goyal, immunogenicity of nicotine vaccines obtained by intradermal delivery with safe R. Brooking-Dixon, H. F. Staats, and S. N. Abraham. 2008. Mast cell activa- adjuvants. Vaccine 31: 159–164. tors: a new class of highly effective vaccine adjuvants. Nat. Med. 14: 536–541. 10. Kashiwagi, S., T. Brauns, J. Gelfand, and M. C. Poznansky. 2014. Laser vaccine 39. St John, A. L., C. Y. Chan, H. F. Staats, K. W. Leong, and S. N. Abraham. 2012. http://www.jimmunol.org/ adjuvants. History, progress, and potential. Hum. Vaccin. Immunother. 10: Synthetic mast-cell granules as adjuvants to promote and polarize immunity in 1892–1907. lymph nodes. Nat. Mater. 11: 250–257. 11. Kashiwagi, S., T. Brauns, and M. C. Poznansky. 2016. Classification of laser 40. Gupta, R. K., B. E. Rost, E. Relyveld, and G. R. Siber. 1995. Adjuvant prop- vaccine adjuvants. J. Vaccines Vaccin. 7: 307. erties of aluminum and calcium compounds. Pharm. Biotechnol. 6: 229–248. 12. Wang, J., D. Shah, X. Chen, R. R. Anderson, and M. X. Wu. 2014. A micro-sterile 41. Wang, L., D. Zhang, and W. Schwarz. 2014. TRPV channels in mast cells as a inflammation array as an adjuvant for influenza vaccines. Nat. Commun. 5: 4447. target for low-level-laser therapy. Cells 3: 662–673. 13. Wang, J., B. Li, and M. X. Wu. 2015. Effective and lesion-free cutaneous in- 42. Freichel, M., J. Almering, and V. Tsvilovskyy. 2012. The role of TRP fluenza vaccination. Proc. Natl. Acad. Sci. USA 112: 5005–5010. in mast cells. Front. Immunol. 3: 150. 14. Terhorst, D., R. Chelbi, C. Wohn, C. Malosse, S. Tamoutounour, A. Jorquera, 43. Suto, H., S. Nakae, M. Kakurai, J. D. Sedgwick, M. Tsai, and S. J. Galli. 2006. M. Bajenoff, M. Dalod, B. Malissen, and S. Henri. 2015. Dynamics and tran- Mast cell-associated TNF promotes dendritic cell migration. J. Immunol. 176: scriptomics of skin dendritic cells and macrophages in an imiquimod-induced, 4102–4112. biphasic mouse model of psoriasis. J. Immunol. 195: 4953–4961. by guest on September 27, 2021 44. Karu, T. I. 2008. Mitochondrial signaling in mammalian cells activated by red 15. Onikienko, S. B., A. B. Zemlyanoy, B. A. Margulis, I. V. Guzhova, and near-IR radiation. Photochem. Photobiol. 84: 1091–1099. M. B. Varlashova, V. S. Gornostaev, N. V. Tikhonova, G. A. Baranov, and 45. Barolet, D., F. Christiaens, and M. R. Hamblin. 2016. Infrared and skin: friend V. V. Lesnichiy. 2007. Diagnostics and correction of the metabolic and immune or foe. J. Photochem. Photobiol. B 155: 78–85. disorders. Interactions of bacterial endotoxins and lipophilic xenobiotics with re- 46. Darvin, M. E., S. Haag, M. Meinke, L. Zastrow, W. Sterry, and J. Lademann. ceptors associated with innate immunity. Donosologiya (St. Petersburg) 1: 32–54. 2010. Radical production by infrared A irradiation in human tissue. Skin 16. Chen, X., P. Kim, B. Farinelli, A. Doukas, S. H. Yun, J. A. Gelfand, Pharmacol. Physiol. 23: 40–46. R. R. Anderson, and M. X. Wu. 2010. A novel laser vaccine adjuvant increases 47. Zastrow, L., N. Groth, F. Klein, D. Kockott, J. Lademann, R. Renneberg, and the motility of antigen presenting cells. PLoS One 5: e13776. L. Ferrero. 2009. The missing link--light-induced (280-1,600 nm) free radical 17. Chen, X., Q. Zeng, and M. X. Wu. 2012. Improved efficacy of dendritic cell- formation in human skin. Skin Pharmacol. Physiol. 22: 31–44. based immunotherapy by cutaneous laser illumination. Clin. Res. 18: 48. Swindle, E. J., and D. D. Metcalfe. 2007. The role of reactive oxygen species 2240–2249. and nitric oxide in mast cell-dependent inflammatory processes. Immunol. Rev. 18. Kimizuka, Y., J. J. Callahan, Z. Huang, K. Morse, W. Katagiri, A. Shigeta, 217: 186–205. R. Bronson, S. Takeuchi, Y. Shimaoka, M. P. K. Chan, et al. 2017. Semicon- 49. Kalesnikoff, J., and S. J. Galli. 2011. Antiinflammatory and immunosuppres- ductor diode laser device adjuvanting intradermal vaccine. Vaccine 35: 2404– sive functions of mast cells. Methods Mol. Biol. 677: 207–220. 2412. 50. Kalyanaraman, B., V. Darley-Usmar, K. J. Davies, P. A. Dennery, H. J. Forman, 19. Kashiwagi, S., J. Yuan, B. Forbes, M. L. Hibert, E. L. Lee, L. Whicher, M. B. Grisham, G. E. Mann, K. Moore, L. J. Roberts, II, and H. Ischiropoulos. C. Goudie, Y. Yang, T. Chen, B. Edelblute, et al. 2013. Near-infrared laser 2012. Measuring reactive oxygen and nitrogen species with fluorescent probes: adjuvant for influenza vaccine. PLoS One 8: e82899. challenges and limitations. Free Radic. Biol. Med. 52: 1–6. 20. Morse, K., Y. Kimizuka, M. P. K. Chan, M. Shibata, Y. Shimaoka, S. Takeuchi, 51. Schmitt, F. J., G. Renger, T. Friedrich, V. D. Kreslavski, B. Forbes, C. Nirschl, B. Li, Y. Zeng, et al. 2017. Near-infrared 1064 nm laser S. K. Zharmukhamedov, D. A. Los, V. V. Kuznetsov, and S. I. Allakhverdiev. modulates migratory dendritic cells to augment the immune response to in- 2014. Reactive oxygen species: re-evaluation of generation, monitoring and tradermal influenza vaccine. J. Immunol. 199: 1319–1332. role in stress-signaling in phototrophic organisms. Biochim. Biophys. Acta 21. Niemz, M. H. 2007. Laser-Tissue Interactions: Fundamentals and Applications 1837: 835–848. (Biological and Medical Physics, Biomedical Engineering). Springer, Berlin. 52. Bolisetty, S., and E. A. Jaimes. 2013. Mitochondria and reactive oxygen spe- 22. Carrasco, E., M. I. Calvo, A. Bla´zquez-Castro, D. Vecchio, A. Zamarro´n, I. J. ´ cies: physiology and pathophysiology. Int. J. Mol. Sci. 14: 6306–6344. D. de Almeida, J. C. Stockert, M. R. Hamblin, A. Juarranz, and J. Espada. 2015. 53. Malissen, B., S. Tamoutounour, and S. Henri. 2014. The origins and functions Photoactivation of ROS production in situ transiently activates cell proliferation of dendritic cells and macrophages in the skin. Nat. Rev. Immunol. 14: 417–428. in mouse skin and in the hair follicle niche promoting hair growth and 54. Clausen, B. E., and P. Stoitzner. 2015. Functional specialization of skin den- wound healing. J. Invest. Dermatol. 135: 2611–2622. dritic cell subsets in regulating T cell responses. Front. Immunol. 6: 534. 23. Macatonia, S. E., S. C. Knight, A. J. Edwards, S. Griffiths, and P. Fryer. 1987. 55. Hildner, K., B. T. Edelson, W. E. Purtha, M. Diamond, H. Matsushita, Localization of antigen on dendritic cells after exposure to the M. Kohyama, B. Calderon, B. U. Schraml, E. R. Unanue, M. S. Diamond, et al. contact sensitizer fluorescein isothiocyanate. Functional and morphological 2008. Batf3 deficiency reveals a critical role for CD8alpha+ dendritic cells in studies. J. Exp. Med. 166: 1654–1667. cytotoxic T cell immunity. Science 322: 1097–1100. 24. Anandasabapathy, N., R. Feder, S. Mollah, S. W. Tse, M. P. Longhi, 56. Martı´nez-Lo´pez, M., S. Iborra, R. Conde-Garrosa, and D. Sancho. 2015. Batf3- S. Mehandru, I. Matos, C. Cheong, D. Ruane, L. Brane, et al. 2014. Classical dependent CD103+ dendritic cells are major producers of IL-12 that drive local Flt3L-dependent dendritic cells control immunity to vaccine. J. Exp. Th1 immunity against Leishmania major infection in mice. Eur. J. Immunol. 45: Med. 211: 1875–1891. 119–129. 25. Mollah, S. A., J. S. Dobrin, R. E. Feder, S. W. Tse, I. G. Matos, C. Cheong, 57. Igya´rto´, B. Z., K. Haley, D. Ortner, A. Bobr, M. Gerami-Nejad, B. T. Edelson, R. M. Steinman, and N. Anandasabapathy. 2014. Flt3L dependence helps define S. M. Zurawski, B. Malissen, G. Zurawski, J. Berman, and D. H. Kaplan. 2011. an uncharacterized subset of murine cutaneous dendritic cells. J. Invest. Skin-resident murine dendritic cell subsets promote distinct and opposing Dermatol. 134: 1265–1275. antigen-specific T helper cell responses. Immunity 35: 260–272. The Journal of Immunology 17

58. Ohl, L., M. Mohaupt, N. Czeloth, G. Hintzen, Z. Kiafard, J. Zwirner, 85. Tanaka, Y., K. Matsuo, S. Yuzuriha, H. Yan, and J. Nakayama. 2010. Non- T. Blankenstein, G. Henning, and R. Fo¨rster. 2004. CCR7 governs skin den- thermal cytocidal effect of infrared irradiation on cultured cancer cells using dritic cell migration under inflammatory and steady-state conditions. Immunity specialized device. Cancer Sci. 101: 1396–1402. 21: 279–288. 86. Tanaka, Y., N. Tatewaki, H. Nishida, T. Eitsuka, N. Ikekawa, and J. Nakayama. 59. Fo¨rster, R., A. Schubel, D. Breitfeld, E. Kremmer, I. Renner-Mu¨ller, E. Wolf, 2012. Non-thermal DNA damage of cancer cells using near-infrared irradiation. and M. Lipp. 1999. CCR7 coordinates the primary immune response by Cancer Sci. 103: 1467–1473. establishing functional microenvironments in secondary lymphoid organs. Cell 87. Zorov, D. B., M. Juhaszova, and S. J. Sollott. 2006. Mitochondrial ROS-induced 99: 23–33. ROSrelease:anupdateandreview.Biochim. Biophys. Acta 1757: 509–517. 60. Khazaei, M., F. Moien-Afshari, and I. Laher. 2008. Vascular endothelial 88. Passarella, S., and T. Karu. 2014. Absorption of monochromatic and narrow function in health and diseases. Pathophysiology 15: 49–67. band radiation in the visible and near IR by both mitochondrial and non- 61. Worbs, T., S. I. Hammerschmidt, and R. Fo¨rster. 2017. Dendritic cell migration mitochondrial photoacceptors results in photobiomodulation. J. Photochem. in health and disease. Nat. Rev. Immunol. 17: 30–48. Photobiol. B 140: 344–358. 62. Oliver, G., and M. Detmar. 2002. The rediscovery of the : old 89. Chelombitko, M. A., A. V. Fedorov, O. P. Ilyinskaya, R. A. Zinovkin, and and new insights into the development and biological function of the lymphatic B. V. Chernyak. 2016. Role of reactive oxygen species in mast cell degranu- vasculature. Genes Dev. 16: 773–783. lation. Biochemistry (Mosc.) 81: 1564–1577. 63. Russo, E., M. Nitschke´, and C. Halin. 2013. Dendritic cell interactions with 90. Kim, H. B., K. Y. Baik, P. H. Choung, and J. H. Chung. 2017. Pulse frequency lymphatic endothelium. Lymphat. Res. Biol. 11: 172–182. dependency of photobiomodulation on the bioenergetic functions of human 64. Kunder, C. A., A. L. St John, and S. N. Abraham. 2011. Mast cell modulation of dental pulp stem cells. Sci. Rep. 7: 15927. the vascular and lymphatic endothelium. Blood 118: 5383–5393. 91. Mohanty, S. K., M. Sharma, and P. K. Gupta. 2006. Generation of ROS in cells 65. Schemann, M., E. M. Kugler, S. Buhner, C. Eastwood, J. Donovan, W. Jiang, on exposure to CW and pulsed near-infrared laser tweezers. Photochem. Pho- and D. Grundy. 2012. The mast cell degranulator compound 48/80 directly tobiol. Sci. 5: 134–139. activates neurons. PLoS One 7: e52104. 92. Keshri, G. K., A. Gupta, A. Yadav, S. K. Sharma, and S. B. Singh. 2016. 66. Theoharides, T. C., K. D. Alysandratos, A. Angelidou, D. A. Delivanis, Photobiomodulation with pulsed and continuous wave near-infrared laser (810 N. Sismanopoulos, B. Zhang, S. Asadi, M. Vasiadi, Z. Weng, A. Miniati, and nm, Al-Ga-as) augments dermal wound healing in immunosuppressed rats. D. Kalogeromitros. 2012. Mast cells and inflammation. Biochim. Biophys. Acta PLoS One 11: e0166705. 1822: 21–33. 93. Hashmi, J. T., Y. Y. Huang, S. K. Sharma, D. B. Kurup, L. De Taboada, Downloaded from 67. Theoharides, T. C., D. Kempuraj, M. Tagen, P. Conti, and D. Kalogeromitros. J. D. Carroll, and M. R. Hamblin. 2010. Effect of pulsing in low-level light 2007. Differential release of mast cell mediators and the pathogenesis of in- therapy. Lasers Surg. Med. 42: 450–466. flammation. Immunol. Rev. 217: 65–78. 94. Cho, S., M. H. Shin, Y. K. Kim, J. E. Seo, Y. M. Lee, C. H. Park, and 68. Randolph, G. J., V. Angeli, and M. A. Swartz. 2005. Dendritic-cell trafficking J. H. Chung. 2009. Effects of infrared radiation and heat on human skin aging to lymph nodes through lymphatic vessels. Nat. Rev. Immunol. 5: 617–628. in vivo. J. Investig. Dermatol. Symp. Proc. 14: 15–19. 69. Angeli, V., and G. J. Randolph. 2006. Inflammation, lymphatic function, and 95. Nedoszytko, B., M. Sokołowska-Wojdyło, K. Ruckemann-Dziurdzinska, dendritic cell migration. Lymphat. Res. Biol. 4: 217–228. J. Roszkiewicz, and R. J. Nowicki. 2014. Chemokines and cytokines network in

70. Johnson, L. A., and D. G. Jackson. 2010. Inflammation-induced secretion of the pathogenesis of the inflammatory skin diseases: atopic dermatitis, psoriasis http://www.jimmunol.org/ CCL21 in lymphatic endothelium is a key regulator of integrin-mediated and skin . Postepy Dermatol. Alergol. 31: 84–91. dendritic cell transmigration. Int. Immunol. 22: 839–849. 96. Kondo, S. 1999. The roles of keratinocyte-derived cytokines in the epidermis 71. Ohshiro, T. 2012. Personal overview of the application of LLLT in severely and their possible responses to UVA-irradiation. J. Investig. Dermatol. Symp. infertile Japanese females. Laser Ther. 21: 97–103. Proc. 4: 177–183. 72. Ohshiro, T. 2012. The proximal priority theory: an updated technique in low 97. Nestle, F. O., P. Di Meglio, J. Z. Qin, and B. J. Nickoloff. 2009. Skin immune level laser therapy with an 830 nm GaAlAs laser. Laser Ther. 21: 275–285. sentinels in health and disease. Nat. Rev. Immunol. 9: 679–691. 73. Huang, Y. Y., A. C. Chen, J. D. Carroll, and M. R. Hamblin. 2009. Biphasic 98. Albanesi, C., S. Madonna, P. Gisondi, and G. Girolomoni. 2018. The interplay dose response in low level light therapy. Dose Response 7: 358–383. between keratinocytes and immune cells in the pathogenesis of psoriasis. Front. 74. Huang, Y. Y., S. K. Sharma, J. Carroll, and M. R. Hamblin. 2011. Biphasic dose Immunol. 9: 1549. response in low level light therapy - an update. Dose Response 9: 602–618. 99. Wang, Z., N. Mascarenhas, L. Eckmann, Y. Miyamoto, X. Sun, T. Kawakami, 75. Karu, T. 2013. Is it time to consider photobiomodulation as a drug equivalent? and A. Di Nardo. 2017. Skin microbiome promotes mast cell maturation by

Photomed. Laser Surg. 31: 189–191. triggering stem cell factor production in keratinocytes. J. Allergy Clin. Immu- by guest on September 27, 2021 76. Karu, T. 2010. Mitochondrial mechanisms of photobiomodulation in con- nol. 139: 1205–1216.e6. text of new data about multiple roles of ATP. Photomed. Laser Surg. 28: 100. Welker, P., J. Grabbe, and B. M. Czarnetzki. 1995. Human keratinocytes release 159–160. mast cell differentiation factors other than stem cell factor. Int. Arch. Allergy 77. Karu, T. I., and S. F. Kolyakov. 2005. Exact action spectra for cellular responses Immunol. 107: 139–141. relevant to phototherapy. Photomed. Laser Surg. 23: 355–361. 101. Algermissen, B., B. Hermes, I. Feldmann-Boeddeker, F. Bauer, and 78. Saenko, Y. V., E. S. Glushchenko, I. O. Zolotovskii, E. Sholokhov, and B. M. Henz. 1999. Mast cell chymase and tryptase during tissue turnover: A. Kurkov. 2016. Mitochondrial dependent oxidative stress in cell culture in- analysis on in vitro mitogenesis of fibroblasts and keratinocytes and alterations duced by laser radiation at 1265 nm. Lasers Med. Sci. 31: 405–413. in cutaneous scars. Exp. Dermatol. 8: 193–198. 79. Sokolovski, S. G., S. A. Zolotovskaya, A. Goltsov, C. Pourreyron, A. P. South, 102. Artuc, M., U. M. Steckelings, A. Gru¨tzkau, A. Smorodchenko, and B. M. Henz. and E. U. Rafailov. 2013. Infrared laser pulse triggers increased singlet oxygen 2002. A long-term coculture model for the study of mast cell-keratinocyte production in tumour cells. Sci. Rep. 3: 3484. interactions. J. Invest. Dermatol. 119: 411–415. 80. Anquez, F., I. El Yazidi-Belkoura, S. Randoux, P. Suret, and E. Courtade. 2012. 103. Mukai, K., M. Tsai, H. Saito, and S. J. Galli. 2018. Mast cells as sources of Cancerous cell death from sensitizer free photoactivation of singlet oxygen. cytokines, chemokines, and growth factors. Immunol. Rev. 282: 121–150. Photochem. Photobiol. 88: 167–174. 104. Tu¨zu¨n, Y., M. Antonov, N. Dolar, and R. Wolf. 2007. Keratinocyte cytokine and 81. Schroeder, P., C. Pohl, C. Calles, C. Marks, S. Wild, and J. Krutmann. 2007. chemokine receptors. Dermatol. Clin. 25: 467–476, vii. Cellular response to infrared radiation involves retrograde mitochondrial sig- 105. Van Loveren, H., S. Kraeuter-Kops, and P. W. Askenase. 1984. Different naling. Free Radic. Biol. Med. 43: 128–135. mechanisms of release of vasoactive amines by mast cells occur in T cell- 82. Piazena, H., W. Pittermann, W. Mu¨ller, K. Jung, D. K. Kelleher, T. Herrling, dependent compared to IgE-dependent cutaneous hypersensitivity responses. P. Meffert, R. Uebelhack, and M. Kietzmann. 2014. Effects of water-filtered Eur. J. Immunol. 14: 40–47. infrared-A and of heat on cell death, inflammation, antioxidative potential and 106. Letourneau, R., X. Pang, G. R. Sant, and T. C. Theoharides. 1996. Intragranular of free radical formation in viable skin--first results. J. Photochem. Photobiol. B activation of bladder mast cells and their association with nerve processes in 138: 347–354. interstitial cystitis. Br.J.Urol.77: 41–54. 83. Schieke, S., H. Stege, V. Ku¨rten, S. Grether-Beck, H. Sies, and J. Krutmann. 107. Dimitriadou, V., M. G. Buzzi, M. A. Moskowitz, and T. C. Theoharides. 1991. 2002. Infrared-A radiation-induced matrix metalloproteinase 1 expression is Trigeminal sensory fiber stimulation induces morphological changes reflecting mediated through extracellular signal-regulated kinase 1/2 activation in human secretion in rat dura mater mast cells. Neuroscience 44: 97–112. dermal fibroblasts. J. Invest. Dermatol. 119: 1323–1329. 108. Dvorak, A. M., R. S. McLeod, A. Onderdonk, R. A. Monahan-Earley, 84. Schroeder, P., J. Lademann, M. E. Darvin,H.Stege,C.Marks,S.Bruhnke, J. B. Cullen, D. A. Antonioli, E. Morgan, J. E. Blair, P. Estrella, R. L. Cisneros, and J. Krutmann. 2008. Infrared radiation-induced matrix metalloproteinase et al. 1992. Ultrastructural evidence for piecemeal and anaphylactic degranu- in human skin: implications for protection. J. Invest. Dermatol. 128: 2491– lation of human gut mucosal mast cells in vivo. Int. Arch. Allergy Immunol. 99: 2497. 74–83.