Human variation in gingival inflammation

Shatha Bamashmous, Georgios A Kotsakis, Kristopher A Kerns, Brian G Leroux, Camille Zenobia, Dandan Chen, Harsh M Trivedi, Jeffrey S McLean, Richard P Darveau, Shatha Bamashmous, Georgios A Kotsakis, Kristopher A Kerns, Brian G Leroux, Camille Zenobia, Dandan Chen, Harsh M Trivedi, Jeffrey S McLean, Richard P Darveau

Abstract

Oral commensal bacteria actively participate with gingival tissue to maintain healthy neutrophil surveillance and normal tissue and bone turnover processes. Disruption of this homeostatic host-bacteria relationship occurs during experimental gingivitis studies where it has been clearly established that increases in the bacterial burden increase gingival inflammation. Here, we show that experimental gingivitis resulted in three unique clinical inflammatory phenotypes (high, low, and slow) and reveal that interleukin-1β, a reported major gingivitis-associated inflammatory mediator, was not associated with clinical gingival inflammation in the slow response group. In addition, significantly higher levels of Streptococcus spp. were also unique to this group. The low clinical response group was characterized by low concentrations of host mediators, despite similar bacterial accumulation and compositional characteristics as the high clinical response group. Neutrophil and bone activation modulators were down-regulated in all response groups, revealing novel tissue and bone protective responses during gingival inflammation. These alterations in chemokine and microbial composition responses during experimental gingivitis reveal a previously uncharacterized variation in the human host response to a disruption in gingival homeostasis. Understanding this human variation in gingival inflammation may facilitate the identification of periodontitis-susceptible individuals. Overall, this study underscores the variability in host responses in the human population arising from variations in host immune profiles (low responders) and microbial community maturation (slow responders) that may impact clinical outcomes in terms of destructive inflammation.

Keywords: chemokine; gingivitis; inflammation; oral microbiome; periodontitis.

Conflict of interest statement

Competing interest statement: This work was funded in part by a Colgate-Palmolive clinical research grant (R.P.D., PI). This funding source had no role in the design of this study and did not have any role during its implementation, analyses, interpretation of the data, or decision to submit results.

Copyright © 2021 the Author(s). Published by PNAS.

Figures

Fig. 1.
Fig. 1.
Experimental gingivitis study design. (A) Experimental cessation of oral hygiene leads to increased plaque biomass and induced gingivitis. Within-subject contralateral teeth with regular oral hygiene served as controls. Baseline was established from day −14 to day 0. Induction of experimental gingival inflammation was carried out from day 0 through day 21. Resolution of experimental gingivitis was observed from day 21 until day 35 when no residual clinical inflammatory activity was detectable (see SI Appendix, Supplemental Materials and Methods). Comprehensive clinical, chemokine and microbiome analysis were conducted to obtain highly granular multiplexed analyses of changes during induction and resolution of inflammation. (B) Percentages of subjects that were clustered into the three clinical response groups (high, low, and slow) based on longitudinal trajectories of clinical parameters (GI, PI, and BOP) (SI Appendix, Fig. S1).
Fig. 2.
Fig. 2.
Differential clinical inflammatory responses and chemokine levels in the three response groups versus controls. (AD) Distributions of clinical parameters comparing test side (no-brushing) versus control side (brushed) for all subjects over time. (EH) Temporal changes in inflammatory-associated clinical measures stratified by response group (high [n = 6], low [n = 6], slow [n = 9]). (IL) Responses among sample-sites in each of three clinical inflammatory response groups (high, low, and slow) before, during, and after resolution of bacterial-induced inflammation. (I) Bacterial load based on total 16S rRNA gene copies; y axis log10-scaled. (J) Neutrophil marker MPO. (K and L) proinflammatory cytokine IL-1β. Boxes represent data and medians ± interquartile ranges (IQR); whiskers and outliers > 1.5 IQR below (above) the 25th (75th) percentile. Trend lines represent mean values across time points. Different letters above bars indicate the significant differences between groups at that time point (a, b, c) (FDR P < 0.05). In AD statistical analyses were performed against the controls by comparing all test samples (prior to responder group identification) and all pooled data from the control teeth data (intraoral control) for just these select clinical parameters and only calculated for the induction phase. In EL the responder groups are shown in relation to the pooled control group. Control samples were never physically pooled; however, the data, for clarity in the figures, was reduced from the total six groups (three test and three control) to four groups (three test and one pooled control). Differences of each group compared to baseline (day 0) are shown above the groups and their significance level indicated by asterisks. Significance levels: *P < 0.05, **P ≤ 0.01, and ***P ≤ 0.001. (L) Comparison of IL-1β levels between groups at peak clinical inflammatory endpoint (day 21) show the subjects in the slow group were significantly different from the high responder group that displayed elevated IL-1β.
Fig. 3.
Fig. 3.
Relationship between chemokine levels and the three clinical response groups. (A) Normalized (row-wise z-score) mean values for chemokine expression by responder groups and controls (all responder groups’ control side data are combined for clarity) by day. The low responder group displays several SDs below the mean of all the groups in most chemokines. (B) Temporal relationships between major neutrophil chemokines IL-8/CXCL8 and MIF with MPO and bacterial load. MIF demonstrated an inverted U-shape distribution during the induction-resolution human experiment that was similar to the temporal changes in microbial load (negative quadratic coefficients; P < 0.01 for both MIF and bacterial load). MPO changes followed the same temporal pattern as MIF (negative quadratic coefficient; P < 0.001), while IL-8/CXCL8 levels demonstrated a U-shape distribution with no association to MPO levels (positive quadratic coefficient; P = 0.09). (CH) Temporal changes in major neutrophil chemokines (MIF, IL-8/CXCL8, and GCP-2/CXCL6) across responder groups compared to each of the responder groups (C, E, G) and between their respective control side sample in the split-mouth design (D, F, H). Boxes represents data and medians ± IQR; whiskers and outliers > 1.5 IQR below (above) the 25th (75th) percentile. Trend lines represent mean values across time points. For D, F, and H, separate controls for each group in were displayed against their respective healthy controls. Differences of each group compared to baseline (day 0) are shown above the groups and their significance level indicated by asterisks. Significance levels: *P < 0.05 and **P ≤ 0.01.
Fig. 4.
Fig. 4.
(A–F) Changes in levels of chemokines involved in bone homeostasis following induction of reversible bone-sparing gingival inflammation. Responses among sample-sites in each of three clinical inflammatory response groups before, during, and after resolution of bacterial-induced inflammation. Boxes represents data and medians ± IQR; whiskers and outliers > 1.5 IQR. Trend lines represent mean values across time points. Different letters above bars indicate the significant differences between groups at that time point (a, b, c) (FDR P < 0.05). Differences of each group compared to baseline (day 0) are shown above the groups and their significance level indicated by asterisks. Significance levels: *P < 0.05, **P ≤ 0.01, and ***P ≤ 0.001.
Fig. 5.
Fig. 5.
Temporal changes in microbial diversity and taxonomy vary by inflammatory responder type. (A) Faith’s phylogenetic diversity by responder group and controls from day −14 to day 35 (n = 334 samples from 21 subjects). Boxplots show median and 25th/75th quartiles; whiskers show inner fences. Lines show mean richness by clinical responder group (high, low, and slow) and controls (within-subject noninflamed gingival sites). (B) Nonmetric multidimensional scaling (NMDS) plots of β-diversity (unweighted Unifrac distance matrices) by responder group and controls from day −14 to day 35 (n = 334 samples from 21 subjects). Tests for significance in β-diversity between groups determined by PERMANOVA. (C) Phylum-level distributions of relative abundance by responder group and controls from day −14 to day 35 (n = 334 samples from 21 subjects). (D and E) Genus-level mean relative abundance by responder group and controls from day −14 to day 35. Linear regression (Loess) shown with 95% confidence bound (SI Appendix, Supplemental Materials and Methods) (n = 334 samples from 21 subjects). (F) Streptococcus mean relative abundance by responder group and controls from day −14 to day 35 (n = 334 samples from 21 subjects). Also shown are center log-transformed (CLR) relative abundances of ASVs taxonomically assigned to S. sanguinis and S. oralis species in high, slow, and low inflammatory response groups from day −14 and day 35. Boxplots show median and lower/upper quartiles; whiskers show inner fences (Materials and Methods) (n = 84 from 21 subjects), (Wilcoxon test; adjusted by FDR) (SI Appendix, Supplemental Materials and Methods and Table S5). Asterisks show FDR-corrected statistical significance levels (FDR **P ≤ 0.01 and ****P ≤ 0.0001).

References

    1. Medzhitov R., Origin and physiological roles of inflammation. Nature 454, 428–435 (2008).
    1. Chen M., Xu H., Parainflammation, chronic inflammation, and age-related macular degeneration. J. Leukoc. Biol. 98, 713–725 (2015).
    1. Fine N., et al. ., Distinct oral neutrophil subsets define health and periodontal disease states. J. Dent. Res. 95, 931–938 (2016).
    1. Hooper L. V., Bacterial contributions to mammalian gut development. Trends Microbiol. 12, 129–134 (2004).
    1. Darveau R. P., Periodontitis: A polymicrobial disruption of host homeostasis. Nat. Rev. Microbiol. 8, 481–490 (2010).
    1. Tonetti M. S., Imboden M. A., Lang N. P., Neutrophil migration into the gingival sulcus is associated with transepithelial gradients of interleukin-8 and ICAM-1. J. Periodontol. 69, 1139–1147 (1998).
    1. Hajishengallis G., Darveau R. P., Curtis M. A., The keystone-pathogen hypothesis. Nat. Rev. Microbiol. 10, 717–725 (2012).
    1. Curtis M. A., Zenobia C., Darveau R. P., The relationship of the oral microbiotia to periodontal health and disease. Cell Host Microbe 10, 302–306 (2011).
    1. Tsukamoto Y., et al. ., Role of the junctional epithelium in periodontal innate defense and homeostasis. J. Periodontal Res. 47, 750–757 (2012).
    1. Zenobia C., et al. ., Commensal bacteria-dependent select expression of CXCL2 contributes to periodontal tissue homeostasis. Cell. Microbiol. 15, 1419–1426 (2013).
    1. Hajishengallis G., et al. ., Low-abundance biofilm species orchestrates inflammatory periodontal disease through the commensal microbiota and complement. Cell Host Microbe 10, 497–506 (2011).
    1. Irie K., Novince C. M., Darveau R. P., Impact of the oral commensal flora on alveolar bone homeostasis. J. Dent. Res. 93, 801–806 (2014).
    1. Darveau R. P., The oral microbial consortium’s interaction with the periodontal innate defense system. DNA Cell Biol. 28, 389–395 (2009).
    1. Löe H., Theilade E., Jensen S. B., Experimental gingivitis in man. J. Periodontol. 36, 177–187 (1965).
    1. Kurgan S., Kantarci A., Molecular basis for immunohistochemical and inflammatory changes during progression of gingivitis to periodontitis. Periodontol. 2000 76, 51–67 (2018).
    1. Zemouri C., et al. ., Resistance and resilience to experimental gingivitis: A systematic scoping review. BMC Oral Health 19, 212 (2019).
    1. Tatakis D. N., Trombelli L., Modulation of clinical expression of plaque-induced gingivitis. I. Background review and rationale. J. Clin. Periodontol. 31, 229–238 (2004).
    1. Trombelli L., et al. ., Experimental gingivitis: Reproducibility of plaque accumulation and gingival inflammation parameters in selected populations during a repeat trial. J. Clin. Periodontol. 35, 955–960 (2008).
    1. Trombelli L., et al. ., Modulation of clinical expression of plaque-induced gingivitis. II. Identification of “high-responder” and “low-responder” subjects. J. Clin. Periodontol. 31, 239–252 (2004).
    1. Löe H., Silness J., Periodontal disease in pregnancy I. Prevalence and severity. Acta Odontol. Scand. 21, 533–551 (1963).
    1. Silness J., Löe H., Periodontal disease in pregnancy II. Correlation between oral hygiene and periodontal condition. Acta Odontol. Scand. 22, 121–135 (1964).
    1. Kurgan S., Kantarci A., Molecular basis for immunohistochemical and inflammatory changes during progression of gingivitis to periodontitis. Periodontol. 2000 76, 51–67 (2018).
    1. Genolini C., Alacoque X., Sentenac M., Arnaud C., kml and kml3d: R packages to cluster longitudinal data. J. Stat. Softw. 65, 1–34 (2015).
    1. Wellappuli N. C., et al. ., Oral and blood neutrophil activation states during experimental gingivitis. JDR Clin. Trans. Res. 3, 65–75 (2018).
    1. Santos L. L., et al. ., Macrophage migration inhibitory factor regulates neutrophil chemotactic responses in inflammatory arthritis in mice. Arthritis Rheum. 63, 960–970 (2011).
    1. Jordan L. A., et al. ., Inhibition of CCL3 abrogated precursor cell fusion and bone erosions in human osteoclast cultures and murine collagen-induced arthritis. Rheumatology (Oxford) 57, 2042–2052 (2018).
    1. Al-Sabbagh M., et al. ., Bone remodeling-associated salivary biomarker MIP-1α distinguishes periodontal disease from health. J. Periodontal Res. 47, 389–395 (2012).
    1. Bhavsar I., Miller C. S., Al-Sabbagh M., “Macrophage inflammatory protein-1 alpha (MIP-1 alpha)/CCL3: As a biomarker” in General Methods in Biomarker Research and Their Applications, Preedy V. R., Patel V. B., Eds. (Springer, 2015), pp. 223–249.
    1. Votta B. J., et al. ., CKbeta-8 [CCL23], a novel CC chemokine, is chemotactic for human osteoclast precursors and is expressed in bone tissues. J. Cell. Physiol. 183, 196–207 (2000).
    1. Stadler A. F., et al. ., Gingival crevicular fluid levels of cytokines/chemokines in chronic periodontitis: A meta-analysis. J. Clin. Periodontol. 43, 727–745 (2016).
    1. Mulholland B. S., Forwood M. R., Morrison N. A., Monocyte chemoattractant protein-1 (MCP-1/CCL2) drives activation of bone remodelling and skeletal metastasis. Curr. Osteoporos. Rep. 17, 538–547 (2019).
    1. Pathak J. L., et al. ., CXCL8 and CCL20 enhance osteoclastogenesis via modulation of cytokine production by human primary osteoblasts. PLoS One 10, e0131041 (2015).
    1. Hosokawa Y., Hosokawa I., Ozaki K., Nakae H., Matsuo T., CXC chemokine ligand 16 in periodontal diseases: Expression in diseased tissues and production by cytokine-stimulated human gingival fibroblasts. Clin. Exp. Immunol. 149, 146–154 (2007).
    1. Bhattacharjee R., et al. ., Efficacy of triphala mouth rinse (aqueous extracts) on dental plaque and gingivitis in children. J. Investig. Clin. Dent. 6, 206–210 (2015).
    1. Hwang J., et al. ., Human CC chemokine CCL23, a ligand for CCR1, induces endothelial cell migration and promotes angiogenesis. Cytokine 30, 254–263 (2005).
    1. Callahan B. J., Sankaran K., Fukuyama J. A., McMurdie P. J., Holmes S. P., Bioconductor workflow for microbiome data analysis: From raw reads to community analyses. F1000 Res. 5, 1492 (2016).
    1. Darveau R. P., Tanner A., Page R. C., The microbial challenge in periodontitis. Periodontol. 2000 14, 12–32 (1997).
    1. Theilade E., Wright W. H., Jensen S. B., Löe H., Experimental gingivitis in man. II. A longitudinal clinical and bacteriological investigation. J. Periodontal Res. 1, 1–13 (1966).
    1. He X., et al. ., Cultivation of a human-associated TM7 phylotype reveals a reduced genome and epibiotic parasitic lifestyle. Proc. Natl. Acad. Sci. U.S.A. 112, 244–249 (2015).
    1. Zee K. Y., Samaranayake L. P., Attström R., An in vivo replica study of microbial colonization in “rapid” and “slow” dental plaque formers. APMIS 108, 113–121 (2000).
    1. Hajishengallis G., The inflammophilic character of the periodontitis-associated microbiota. Mol. Oral Microbiol. 29, 248–257 (2014).
    1. Herrero E. R., et al. ., Antimicrobial effects of commensal oral species are regulated by environmental factors. J. Dent. 47, 23–33 (2016).
    1. Zhang J., Kashket S., Lingström P., Evidence for the early onset of gingival inflammation following short-term plaque accumulation. J. Clin. Periodontol. 29, 1082–1085 (2002).
    1. Offenbacher S., et al. ., Changes in gingival crevicular fluid inflammatory mediator levels during the induction and resolution of experimental gingivitis in humans. J. Clin. Periodontol. 37, 324–333 (2010).
    1. Scapoli C., Tatakis D. N., Mamolini E., Trombelli L., Modulation of clinical expression of plaque-induced gingivitis: Interleukin-1 gene cluster polymorphisms. J. Periodontol. 76, 49–56 (2005).
    1. Trombelli L., et al. ., Interleukin-1 beta levels in gingival crevicular fluid and serum under naturally occurring and experimentally induced gingivitis. J. Clin. Periodontol. 37, 697–704 (2010).
    1. Deinzer R., Weik U., Kolb-Bachofen V., Herforth A., Comparison of experimental gingivitis with persistent gingivitis: Differences in clinical parameters and cytokine concentrations. J. Periodontal Res. 42, 318–324 (2007).
    1. Han K. H., et al. ., Vascular expression of the chemokine CX3CL1 promotes osteoclast recruitment and exacerbates bone resorption in an irradiated murine model. Bone 61, 91–101 (2014).
    1. Greer A., et al. ., Site-specific neutrophil migration and CXCL2 expression in periodontal tissue. J. Dent. Res. 95, 946–952 (2016).
    1. Scott D. A., Krauss J., Neutrophils in periodontal inflammation. Front. Oral Biol. 15, 56–83 (2012).
    1. Pancer B. A., et al. ., Effects of triclosan on host response and microbial biomarkers during experimental gingivitis. J. Clin. Periodontol. 43, 435–444 (2016).
    1. Nonnenmacher C., et al. ., Effect of age on gingival crevicular fluid concentrations of MIF and PGE2. J. Dent. Res. 88, 639–643 (2009).
    1. Bernhagen J., et al. ., MIF is a noncognate ligand of CXC chemokine receptors in inflammatory and atherogenic cell recruitment. Nat. Med. 13, 587–596 (2007).
    1. Madeira M. F., et al. ., MIF induces osteoclast differentiation and contributes to progression of periodontal disease in mice. Microbes Infect. 14, 198–206 (2012).
    1. Rajarathnam K., Schnoor M., Richardson R. M., Rajagopal S., How do chemokines navigate neutrophils to the target site: Dissecting the structural mechanisms and signaling pathways. Cell. Signal. 54, 69–80 (2019).
    1. Cao C. F., Smith Q. T., Crevicular fluid myeloperoxidase at healthy, gingivitis and periodontitis sites. J. Clin. Periodontol. 16, 17–20 (1989).
    1. An J. Y., et al. ., Rapamycin rejuvenates oral health in aging mice. eLife 9, 9 (2020).
    1. Chang A. M., et al. ., Toll-like receptor-2 and -4 responses regulate neutrophil infiltration into the junctional epithelium and significantly contribute to the composition of the oral microbiota. J. Periodontol. 90, 1202–1212 (2019).
    1. Bolyen E., et al. ., QIIME 2: Reproducible, interactive, scalable, and extensible microbiome data science. PeerJ [Preprint] (2018). (Accessed 18 June 2021).
    1. Callahan B. J., et al. ., DADA2: High-resolution sample inference from Illumina amplicon data. Nat. Methods 13, 581–583 (2016).
    1. Chen T., et al. ., The human oral microbiome database: A web accessible resource for investigating oral microbe taxonomic and genomic information. Database (Oxford) 2010, baq013 (2010).
    1. Dewhirst F. E., et al. ., The human oral microbiome. J. Bacteriol. 192, 5002–5017 (2010).
    1. McMurdie P. J., Holmes S., phyloseq: An R package for reproducible interactive analysis and graphics of microbiome census data. PLoS One 8, e61217 (2013).
    1. Genolini C., et al. ., KmL3D: A non-parametric algorithm for clustering joint trajectories. Comput. Methods Bioprograms Biomed. 109, 104–111 (2013).
    1. Bates D., Mächler M., Bolker B., Walker S., Fitting linear mixed-effects models using lme4. J. Stat. Softw. 67, 1–48 (2015).
    1. Benjamini Y., Hochberg Y., Controlling the false discovery rate: A practical and powerful approach to multiple testing. J. R. Stat. Soc. B 57, 289–300 (1995).
    1. Lenth R., Emmeans: Estimated marginal means, aka least-squares means. R Package Version 1 (2018). . Accessed 18 June 2021.
    1. Bamashmous S., Kerns K. A., Darveau R. P., McLean J. S., Human variation in gingival inflammation (HVGI) study. NCBI Sequence Read Archive (SRA). . Deposited 25 March 2020.
    1. Bamashmous S., et al. ., Human variation in gingival inflammation (HVGI) study. Github. . Deposited 8 July 2020.

Source: PubMed

3
구독하다