For decades research has centered on identifying the ideal balanced skin microbiome that prevents disease and on developing therapeutics to foster this balance. However, this single idealized balance may not exist. The skin microbiome changes across the lifespan. This is reflected in the dynamic shifts of the skin microbiome's diverse, inter-connected community of microorganisms with age. While there are core skin microbial taxa, the precise community composition for any individual person is determined by local skin physiology, genetics, microbe–host interactions, and microbe–microbe interactions. As a key interface with the environment, the skin surface and its appendages are also constantly exchanging microbes with close personal contacts and the environment. Hormone fluctuations and immune system maturation also drive age-dependent changes in skin physiology that support different microbial community structures over time. Here, we review recent insights into the factors that shape the skin microbiome throughout life. Collectively, the works summarized within this review highlight how, depending on where we are in lifespan, our skin supports robust microbial communities, while still maintaining microbial features unique to us. This review will also highlight how disruptions to this dynamic microbial balance can influence risk for dermatological diseases as well as impact lifelong health.

The skin harbors complex microecosystems of bacteria, fungi, and viruses, each with distinct adaptions to survive on the skin. However, there is no single definition of a balanced skin microbiome. For any individual the microbial balance is dynamic, maturing with us as we grow and navigate through the environments around us. Across our lifespan, the stability and function of skin microbial communities are driven both by interactions with the host and between microorganisms.

Human skin varies in its physical characteristics across body sites [1–4], ranging from oily / sebaceous to moist or dry, resulting in distinct microenvironments that promote unique microbial populations. Sebaceous sites (e.g. face, chest, and back) have a high density of hair follicles and sebaceous glands. The lipid rich sebum produced by these glands promotes colonization by lipophilic taxa, primarily Cutibacterium bacteria and Malassezia fungi [4–6]. Moist sites (e.g. elbow crease, axilla, and groin) have high concentrations of apocrine sweat glands and are dominated by Staphylococcus spp. and Corynebacterium spp.[4]. In contrast with other sites, dry sites (e.g. forearm, abdomen, and palms) have the lowest abundance yet greatest microbial diversity, with significant populations of Cutibacterium, Corynebacterium, and Streptococcus species [2]. Thus, at a microenvironmental level, the balance of skin microbial communities is partially dictated by these physiologic and abiotic features of the skin niche.

The skin is an immunologically rich organ. To maintain the skin barrier tissue-resident immune cells collectively sample and respond to microbial products [4], produce antimicrobial peptides (AMPs) [7], and ultimately prevent penetration of skin microbes into deeper skin layers or open wounds. A hallmark of a healthy equilibrium among skin microbial communities is the maintenance of skin barrier integrity. This is accomplished by promoting skin cell maturation [8,9], training the immune system [10–12], and preventing pathogen overgrowth through niche exclusion and competitive interactions such as the production of antimicrobials [13–16]. Disruption of this equilibrium is characterized by the overgrowth of some bacterial species, such as Staphylococcus aureus, and an overall loss of community diversity [2]. This dysbiosis can lead to impaired wound healing, increased inflammation, and greater risk for infection [2,17–19].

Finally, microbe–microbe interactions within the skin microbiome can drive overall community structure. The three prominent skin taxa, Cutibacterium acnes, Corynebacterium spp., and coagulase-negative staphylococci (CoNS) are known to mediate other microbial taxa in the microenvironment and have been extensively reviewed elsewhere [5,20]. Key interactions and recent studies are summarized in Table 1.

Table 1
Key microbe–microbe interactions on the skin
Microbe Producing the MoleculeAnti-microbial molecule (Type)Inhibited MicrobesProposed MechanismSource
Cutibacterium acnes Acnecin (peptide) non-anecin producing C. acnes Exact mechanism Unknown. Proposed to help C. acnes phylotypes maintain dominance within a pore. [105,164
Corynebacterium 
Cutimycin (thiopeptide) MRSA Unknown [14
Staphylococcus epidermidis 
Propionic acid (SCFA) MRSA Lowers local pH. This can limit the growth of several pathogens while not significantly affecting growth of skin commensals. [165-168
Escherichia coli 
Candida albicans 
Propionic acid (SCFA) Staphylococcus epidermidis Inhibit bacterial biofilm formation. These SCFA also influence melanocyte, keratinocyte, and sebocyte gene expression, and modulate host inflammation. [114, 169-173
Isobutyric acid (SCFA) 
Isovaleric acid (SCFA) 
Corynebacterium accolens Unidentified protein Staphylococcus aureus Inhibits biofilm formation [13
MRSA 
LipS1 (lipase) Streptococcus pneumoniae Metabolizes host lipids into free fatty acids that inhibit bacterial growth [174,175
Unidentified secreted protein(s) Staphylococcus aureus Inhibits the agr quorum sensing system and prevents agr-dependent virulence factor expression [176
Corynebacterium amycolatum Unidentified secreted protein(s) Staphylococcus aureus Inhibits the agr quorum sensing system and prevents agr-dependent virulence factor expression [176
Corynebacterium striatum Unidentified secreted protein(s) Staphylococcus aureus Inhibits the agr quorum sensing system and prevents agr-dependent virulence factor expression [176
Corynebacterium pseudodiptheriticum Unidentified secreted protein(s) Staphylococcus aureus Inhibits the agr quorum sensing system and prevents agr-dependent virulence factor expression [176
Unidentified secreted factor Staphylococcus aureus Bactericidal against S. aureus when it expresses agr-dependent virulence factors [177
MRSA 
Staphylococcus capitus Capidermicin (AMP) Lactococcus lactis Forms pores in membranes [178
Micrococcus leuteus 
Staphylococcus aureus 
Staphylococcus intermedius 
Staphylococcus pseudointermedis 
Unidentified bacteriocin(s) Listeria monocytogenes Unknown [16
Staphylococcus aureus 
MRSA 
Streptococcus alagactiae 
Streptococcus Bovis 
PSM-beta 1 to PSM-beta 6 Micrococcus leuteus Induces cell lysis [179
PSM 1 to PSM 4 Cutibacterium acnes Act synergistically for targeted killing [180
Staphylococcus caprae Unidentified autoinducing peptide Staphylococcus aureus Inhibits the agr quorum sensing system and prevents agr-dependent virulence factor expression [181
MRSA 
Staphylococcus epidermidis Esp (serine protease) Staphylococcus aureus Inhibits biofilm formation and destructs biofilms [182
Unidentified AMP Staphylococcus aureus Targeted killing [183
Staphylococcus hominis Unidentified AMP Staphylococcus aureus Targeted killing [183
Hominicin (AMP) VRSA Unknown [184
Unidentified bacteriocin(s) Listeria monocytogenes Unknown [16
Streptococcus alagactiae 
Streptococcus Bovis 
Staphylococcus mutans Mutacin 1140 (AMP) VRSA Unknown [184
Staphylococcus lugdunensis Lugdunin (thiazolidine-containing cyclic peptide, AMP) VRSA Unknown [184
Lugdunin (thiazolidine-containing cyclic peptide, AMP) Enterococcus faecium Direct killing [185
Enterococcus faecalis 
Listeria Monocytogenes 
Staphylococcus aureus 
MRSA 
Streptococcus pneumoniae 
Lugdunin (thiazolidine-containing cyclic peptide, AMP) Staphylococcus aureus Direct killing and amplification of innate immune responses [186
Staphylococcus simulans AIP-I to AIP-III Staphylococcus aureus & MRSA Inhibits the agr quorum sensing system and prevents agr-dependent virulence factor expression [187
Staphylococcus warneri AIP-I to AIP-II Staphylococcus aureus & MRSA Inhibits the agr quorum sensing system and prevents agr-dependent virulence factor expression [188
Unidentified bacteriocin(s) Staphylococcus aureus Unknown [16
Streptococcus alagactiae 
Microbe Producing the MoleculeAnti-microbial molecule (Type)Inhibited MicrobesProposed MechanismSource
Cutibacterium acnes Acnecin (peptide) non-anecin producing C. acnes Exact mechanism Unknown. Proposed to help C. acnes phylotypes maintain dominance within a pore. [105,164
Corynebacterium 
Cutimycin (thiopeptide) MRSA Unknown [14
Staphylococcus epidermidis 
Propionic acid (SCFA) MRSA Lowers local pH. This can limit the growth of several pathogens while not significantly affecting growth of skin commensals. [165-168
Escherichia coli 
Candida albicans 
Propionic acid (SCFA) Staphylococcus epidermidis Inhibit bacterial biofilm formation. These SCFA also influence melanocyte, keratinocyte, and sebocyte gene expression, and modulate host inflammation. [114, 169-173
Isobutyric acid (SCFA) 
Isovaleric acid (SCFA) 
Corynebacterium accolens Unidentified protein Staphylococcus aureus Inhibits biofilm formation [13
MRSA 
LipS1 (lipase) Streptococcus pneumoniae Metabolizes host lipids into free fatty acids that inhibit bacterial growth [174,175
Unidentified secreted protein(s) Staphylococcus aureus Inhibits the agr quorum sensing system and prevents agr-dependent virulence factor expression [176
Corynebacterium amycolatum Unidentified secreted protein(s) Staphylococcus aureus Inhibits the agr quorum sensing system and prevents agr-dependent virulence factor expression [176
Corynebacterium striatum Unidentified secreted protein(s) Staphylococcus aureus Inhibits the agr quorum sensing system and prevents agr-dependent virulence factor expression [176
Corynebacterium pseudodiptheriticum Unidentified secreted protein(s) Staphylococcus aureus Inhibits the agr quorum sensing system and prevents agr-dependent virulence factor expression [176
Unidentified secreted factor Staphylococcus aureus Bactericidal against S. aureus when it expresses agr-dependent virulence factors [177
MRSA 
Staphylococcus capitus Capidermicin (AMP) Lactococcus lactis Forms pores in membranes [178
Micrococcus leuteus 
Staphylococcus aureus 
Staphylococcus intermedius 
Staphylococcus pseudointermedis 
Unidentified bacteriocin(s) Listeria monocytogenes Unknown [16
Staphylococcus aureus 
MRSA 
Streptococcus alagactiae 
Streptococcus Bovis 
PSM-beta 1 to PSM-beta 6 Micrococcus leuteus Induces cell lysis [179
PSM 1 to PSM 4 Cutibacterium acnes Act synergistically for targeted killing [180
Staphylococcus caprae Unidentified autoinducing peptide Staphylococcus aureus Inhibits the agr quorum sensing system and prevents agr-dependent virulence factor expression [181
MRSA 
Staphylococcus epidermidis Esp (serine protease) Staphylococcus aureus Inhibits biofilm formation and destructs biofilms [182
Unidentified AMP Staphylococcus aureus Targeted killing [183
Staphylococcus hominis Unidentified AMP Staphylococcus aureus Targeted killing [183
Hominicin (AMP) VRSA Unknown [184
Unidentified bacteriocin(s) Listeria monocytogenes Unknown [16
Streptococcus alagactiae 
Streptococcus Bovis 
Staphylococcus mutans Mutacin 1140 (AMP) VRSA Unknown [184
Staphylococcus lugdunensis Lugdunin (thiazolidine-containing cyclic peptide, AMP) VRSA Unknown [184
Lugdunin (thiazolidine-containing cyclic peptide, AMP) Enterococcus faecium Direct killing [185
Enterococcus faecalis 
Listeria Monocytogenes 
Staphylococcus aureus 
MRSA 
Streptococcus pneumoniae 
Lugdunin (thiazolidine-containing cyclic peptide, AMP) Staphylococcus aureus Direct killing and amplification of innate immune responses [186
Staphylococcus simulans AIP-I to AIP-III Staphylococcus aureus & MRSA Inhibits the agr quorum sensing system and prevents agr-dependent virulence factor expression [187
Staphylococcus warneri AIP-I to AIP-II Staphylococcus aureus & MRSA Inhibits the agr quorum sensing system and prevents agr-dependent virulence factor expression [188
Unidentified bacteriocin(s) Staphylococcus aureus Unknown [16
Streptococcus alagactiae 

Summary table of prominent and recently identified antimicrobial molecules produced by major skin bacterial taxa to inhibit other microbial community members and/or potential pathogens. The table is organized by the taxa that produces the antimicrobial molecule(s). Where possible the molecule's proposed mechanism of action is included.

Abbreviations: AIP, Autoinducing peptide; AMP, Antimicrobial peptide; MRSA, Methicillin resistant Staphylococcus aureus; MSSA, Methicillin sensitive Staphylococcus aureus; PSM, phenol-soluble modulins; SCFA, Short chain fatty acids; VRSA, Vancomycin resistant Staphylococcus aureus.

Homeostatic skin microbial community structures shift as we age and navigate the environment (Figure 1). Throughout a lifetime the skin's physiology changes as the cutaneous immune systems matures and hormones drive sweat and sebum gland development, which impacts the availability of key nutrients. As a direct interface with the environment, the skin also continuously shares microbes with the places and people around us. Below we summarize the shifts in the skin microbiome over the human lifespan and highlight where disruptions to the skin microbiome at critical age-associated stages influence the risk for disease development.

The dynamic balance of the skin and its microbiome over the lifespan.

Figure 1.
The dynamic balance of the skin and its microbiome over the lifespan.

Over a lifetime the skin's physiology changes as an individual's cutaneous immune systems matures and hormones drive sweat and sebum gland development. These changes are associated with shifts in the relative abundance of prominent skin microbial taxa and shifts in the overall microbial community diversity. Microbiome data displays the average relative abundance of the top ten microbial taxa for each group as assessed by high-throughput sequencing of the bacterial 16S ribosomal RNA gene. Taxa with a relative abundance >20% in at least one group are bolded. Groups include newborns born either through vaginal delivery or cesarian section [22] as well as dry, moist, and sebaceous sites for infants (1 year old) [40], children (5 years old) [40], adolescents (Tanner Stage III) [46], adults (20–40 years old) [3], and the elderly (60 and older) [137]. Since sexual differences in skin microbial composition become more pronounced over the course of puberty [46], relative abundance plots for adolescent, adult, and elderly males and females are displayed. Inner circles represent relative microbial diversity, sebum production, sweat production, surface pH, skin integrity, and immune function throughout life [33,46,87,144,152,153].

Figure 1.
The dynamic balance of the skin and its microbiome over the lifespan.

Over a lifetime the skin's physiology changes as an individual's cutaneous immune systems matures and hormones drive sweat and sebum gland development. These changes are associated with shifts in the relative abundance of prominent skin microbial taxa and shifts in the overall microbial community diversity. Microbiome data displays the average relative abundance of the top ten microbial taxa for each group as assessed by high-throughput sequencing of the bacterial 16S ribosomal RNA gene. Taxa with a relative abundance >20% in at least one group are bolded. Groups include newborns born either through vaginal delivery or cesarian section [22] as well as dry, moist, and sebaceous sites for infants (1 year old) [40], children (5 years old) [40], adolescents (Tanner Stage III) [46], adults (20–40 years old) [3], and the elderly (60 and older) [137]. Since sexual differences in skin microbial composition become more pronounced over the course of puberty [46], relative abundance plots for adolescent, adult, and elderly males and females are displayed. Inner circles represent relative microbial diversity, sebum production, sweat production, surface pH, skin integrity, and immune function throughout life [33,46,87,144,152,153].

Close modal

Birth

The skin's first substantial introduction to microbes occurs at birth, where the skin is quickly colonized from the immediate environment. Mode of delivery has been shown to influence the initial community composition [21–23]. For example, the skin microbiome of vaginally delivered newborns is dominated by vaginal-associated flora, primarily Lactobacillus and Prevotella, and contains a higher abundance of Candida albicans [21,22,24]. Newborns delivered via cesarean section have microbiomes containing maternal skin-associated microbes, including Staphylococcus, Streptococcus, Corynebacterium, and Cutibacterium. Cesarian-delivered newborns intentionally exposed to their mother's vaginal flora at birth have microbiomes containing both skin- and vaginal-associated flora [25]. Although these initial communities are transient [22], the order and timing of species colonization influences how strains subsequently interact with one another [26,27]. These priority effects can shape future community structures and have long term implications for the skin, its microbiome, and overall health.

Infancy and childhood: initial microbial exposures

Skin microbial compositions gradually shift throughout infancy and childhood [22]. These communities are shaped by the skin's functional maturation, the microbiomes of close care givers, and their environment [28–30]. Particularly in infancy initial microbial exposures prime immune development [26,31] and strengthen the skin barrier through promoting proper keratinocyte differentiation and epidermal repair [8].

Newborn and infant skin has a greater water content, higher pH, suppressed sebum production, faster epidermal turnover, and greater antimicrobial properties [32,33]. Within hours after birth the skin surface begins to acidify [29] and initially homogenous microbial communities begin to diverge into body site-specific communities [26,34]. Continued site-specific skin maturation promotes reorganization of the infant microbiota across body habitats over the first few months [22]. Within 3–6 months, associations between microbial taxa and skin metabolic function (e.g. lipid production and pH) are established [35]. Compared with adults, reduced sebum production in early life is associated with lower Corynebacterium, Cutibacterium and Malassezia abundance, increased staphylococci and streptococci, and a mycobiome dominated by Candida species [22,24,34,36–39]. As children age, skin further acidifies and produces more sebum lipids, which promotes a gradual decline in acid-sensitive streptococci and increase in overall community diversity [32,33,40].

Commensal colonization also stimulates immune cell maturation, particularly regulatory T-cell (Treg) localization to developing hair follicles shortly after birth [41–43]. Specific training of commensal antigen-specific Tregs is required for establishing immune tolerance [41] that prevents inflammation against this commensal later in life [10]. However early exposure does not guarantee future tolerance. The simultaneous recognition of multiple pathogen-associated molecules and toxins (e.g. S. aureus alpha-toxin) during early colonization events limits pathogen-specific Treg formation [44]. Without microbial-specific Tregs, S. aureus exposures in later life leads to inflammation.

In early infancy, the skin microbiome of close caregivers further contributes to shaping the infant skin microbiome. At 6 weeks of age, infant and maternal skin microbiomes display very similar community structures [22], and throughout childhood the skin will continue to harbor distinct microbial taxa from caregivers [22,40,45]. However with age, older infants have greater skin microbial diversity and more microbes derived from their rural or urban environments [30] and the similarity between a mother and infant microbiome gradually declines [30,45]. Cutaneous fungal populations are also more diverse in childhood with greater interpersonal variation [37,46,47]. These early life variations and interventions in environmental microbial exposures influence the establishment of the skin microbiome and can modulate lifelong immune responses [48–51]. Disruption to the establishment of this equilibrium [26,52] is associated with greater inflammation and can increase a child's risk for atopic dermatitis and allergy development.

Microbial imbalance in pediatric atopic dermatitis

Atopic dermatitis (eczema; AD) is an inflammatory skin disorder, characterized by dry, inflamed, itchy skin patches. The typical age of onset is between 3–6 months old and it affects roughly 20% of infants and children, 15% of adolescents, and 10% of adults [53,54]. This early age of onset and evidence of skin microbiome dysbiosis preceding AD onset [52], underscore how early life microbial exposures may influence AD risk. Additionally, many children grow out of AD and experience partial or full resolution of symptoms as they enter adolescence or adulthood [54]. This showcases the important influence of the maturation of skin microbiome in AD's natural course.

Skin barrier dysfunction and immune imbalance contribute to AD's pathogenesis. In AD patients, both affected and non-affected areas of skin display increased permeability [55], reduced water retention, high pH [56], and altered lipid composition [57–59]. Subsequently external microorganisms, particularly Staphylococcus aureus, can penetrate deeper skin layers. Stressed keratinocytes and microbial antigens trigger type-2 helper T-cell-driven immune responses [60,61], which further exacerbates barrier defects through down-regulation of filaggrin [62,63], disruption of tight junctions, and reduction in stratum corneum lipids [59]. High levels of type-2 cytokines also reduce antimicrobial peptide production, further increasing the skin's susceptivity to S. aureus colonization [63]. Thus, a perpetual cycle of poor barrier function, increased microbial and irritant penetration into deeper skin layers, and increased inflammation then itching and further skin damage is developed.

Elevated S. aureus abundance can precede AD development [26,52] and several S. aureus virulence factors propagate AD related epidermal damage and inflammation [64–66]. Additional microbiome changes in AD include a reduction in Cutibacterium acnes, Corynebacterium, Dermacoccus, Micrococcus, and CoNS and an increase in Streptococcus and some Malassezia species [58,61,67–72]. These microbial shifts appear to be temporal, with a loss of community diversity and greater S. aureus dominance preceding and during AD flares followed by a gradual return to baseline following resolution [73,74]. Finally, early colonization with commensal CoNS (e.g. S. epidermidis and S. hominis) [26] and Gram-negative bacteria reduces the risk of developing AD [75].

Current AD treatment revolves around eliminating exacerbating factors (e.g. contact with allergens and strong soaps), maintaining skin moisture with emollients and reducing skin inflammation through topical corticosteroids or calcineurin inhibiters [54,76]. Many of these interventions effectively resolve the skin dysbiosis [68,73,74,77–79]. Investigations into potential therapies are increasingly centered around modulating the microbiome with oral prebiotics, probiotics and symbiotics [80,81], topical emollients containing probiotics [82], or microbial transplantation with commensal skin bacteria [75,83–85]. One promising avenue is harnessing CoNS anti-S. aureus activity. The transplantation of S. epidermidis and S. hominis can diminish S. aureus virulence, prevent epithelial damage, and reduce inflammation in murine models and clinical trials of atopic dermatitis (AD) patients [84,86].

Adolescence and puberty: hormonally driven changes

Puberty marks the next major shift in our skin microbial communities. The hormones that drive physical and sexual development also directly promote structural and functional changes in the skin such as sebum and apocrine sweat production. This leads to subsequent shifts in microbial composition (Figure 2) [87]. Due to interpersonal variability in puberty onset and progression, the Tanner staging system measures the degree of sexual maturation [88,89]. Tanner stage I corresponds to the pre-pubertal stage and stage V corresponds to adult sexual characteristics. Both cross-sectional and longitudinal studies demonstrate clear shifts in skin microbiome composition across Tanner stages [36,37,46]. Children at stage I have higher relative abundances of Streptococcus, Bacteroidota and Pseudomonadota, as well as higher bacterial and fungal diversity compared with young adults at stage V [36,46]. Furthermore, the young adult skin microbiome is dominated by lipophilic microbes, including Corynebacterium, Cutibacterium acnes, and Malassezia [36,37,46], a composition associated with higher sebum production and serum concentrations of hormones that promote sebum production [46].

Differences in the skin, microbiome, and body odor production in early and late puberty.

Figure 2.
Differences in the skin, microbiome, and body odor production in early and late puberty.

In childhood and early puberty (Tanner Stages I to II) the skin microbiome is highly diverse and body odor is associated with CoNS (e.g., S. epidermidis and S. hominis) production of volatile fatty acids (e.g., propionic, acetic, and isovaleric acid; sour odors) and sulfur (rotten-egg odor) [92]. As puberty advances, steroid hormones promote sebaceous and apocrine sweat gland development [90,91], modify the types of lipids present in sebum [90,97], and enhance the skin barrier [94–96]. In later puberty (Tanner Stages IV to V), increased lipid production and altered lipid content is associated with a skin microbiome dominated by lipophilic taxa [46]. While breakdown of sweat and sebum components into volatile fatty acids still occurs, body odor in young adults becomes more associated with Corynebacterium spp. metabolism of sebum and sweat components into sulfanylalkanols (e.g., 3SH and 3M3SH; oniony odors), and volatile organic compounds (e.g., 3H3MHA; cumin like odors) [92,98–101]. BCAA: Branched chain amino acids; CGSC: Cystine-Glycine-S-conjugate; CoNS: Coagulase negative Staphylococcus spp.; GC: Glutaminyl-conjugate; 3H3MHA: 3-hydroxy-3-methylhexanoic acid; 3M3SH: 3-methyl-3-sullanylhexanol; 3SH: 3-sulfanylhexanol.

Figure 2.
Differences in the skin, microbiome, and body odor production in early and late puberty.

In childhood and early puberty (Tanner Stages I to II) the skin microbiome is highly diverse and body odor is associated with CoNS (e.g., S. epidermidis and S. hominis) production of volatile fatty acids (e.g., propionic, acetic, and isovaleric acid; sour odors) and sulfur (rotten-egg odor) [92]. As puberty advances, steroid hormones promote sebaceous and apocrine sweat gland development [90,91], modify the types of lipids present in sebum [90,97], and enhance the skin barrier [94–96]. In later puberty (Tanner Stages IV to V), increased lipid production and altered lipid content is associated with a skin microbiome dominated by lipophilic taxa [46]. While breakdown of sweat and sebum components into volatile fatty acids still occurs, body odor in young adults becomes more associated with Corynebacterium spp. metabolism of sebum and sweat components into sulfanylalkanols (e.g., 3SH and 3M3SH; oniony odors), and volatile organic compounds (e.g., 3H3MHA; cumin like odors) [92,98–101]. BCAA: Branched chain amino acids; CGSC: Cystine-Glycine-S-conjugate; CoNS: Coagulase negative Staphylococcus spp.; GC: Glutaminyl-conjugate; 3H3MHA: 3-hydroxy-3-methylhexanoic acid; 3M3SH: 3-methyl-3-sullanylhexanol; 3SH: 3-sulfanylhexanol.

Close modal

Puberty is driven by androgens (e.g. dehydroepiandrosterone and testosterone), estrogen, and progesterone [87]. On the skin androgens promote pubic and axillary hair growth and sex specific hair patterns [87], sebaceous gland development and increased sebum production [90,91], and apocrine sweat gland development with a subsequent increase in body odor [92,93]. Estrogen and progesterone enhance the skin barrier and promote wound healing through encouraging collagen synthesis and stimulating keratinocyte proliferation [94–96]. Puberty-driving hormones also modulate immune function in both pro- and anti-inflammatory ways [96]. These hormonally driven changes in skin physiology lead to changes in the skin microbial communities [46]. For instance, androgen-driven sebaceous gland development [91] and estrogen modification of the lipid types at the skin surface [97] promote the survival of lipophilic species, particularly C. acnes and Malassezia restricta [46].

Although apocrine gland secretions are initially odorless, microbial metabolism of their components (e.g. branched chain amino acids, fatty acids, and glycerol) produces odoriferous compounds (Figure 2) [92,98]. Body odor in children and young-teens is associated with microbial production of isovaleric and acetic acid (sour odor) and sulfur (rotten-egg odor) [92]. In young adults body odor, particularly from the axilla [92,98,99], is associated with corynebacterial breakdown of sebum into volatile fatty acids (cheesy odor) and sulfanylalkanols (oniony odors) [99–101].

While all puberty-driving hormones surge in both sexes, the heightened role of androgens in males and estrogens plus progesterone in females foster sexual differences in maturation. As such, sexual differences in skin microbial composition become more pronounced as puberty advances [46]. Microbial communities of females become less diverse and have greater prevalence of Cutibacterium with increasing Tanner stage [46]. Skin microbiomes of males display more inter-individual differences and greater diversity, with higher relative abundance of Corynebacterium, Staphylococcus, Streptococcus, and Haemophilus [46]. Other hormonally dependent skin physiological (lower pH, increased skin thickness) and immunological changes during puberty likely also account for many maturation-associated skin microbial shifts, but their influence is less defined [36,46].

Collectively, puberty marks major shifts in the skin microbial community structure, metabolic complexity, and function. Microbiome imbalance during this process can promote several microbe-associated skin disorders including acne vulgaris.

Cutibacterium acnes phylotype imbalance in acne vulgaris

Several microbe-associated skin disorders often begin during puberty, including acne vulgaris hidradenitis suppurativa, and psoriasis, highlighting the influence of puberty-driven skin and microbial shifts in disease pathogenesis [46]. Here we highlight acne vulgaris (acne), due to the strong support for a microbiome-driven association. Although there are microbial disturbances associated with hidradenitis suppurativa and psoriasis, no single microbe or consistent pattern of dysbiosis has been implicated [61,102,103]. Further research into the roles of the microbiome in these conditions are needed.

Acne affects ∼85% of adolescents and young adults (12–25 years old) [104]. Mild acne is characterized by clogged hair follicles while severe cases present with inflammatory, painful papules, pustules, nodules or cysts. Like how individuals experience resolution of AD as they enter adolescence, acne tends to resolve as individuals mature out of adolescence and enter adulthood. This further highlights how the natural age associated evolution of microbial community structures influence the course of a disease.

Our understanding of acne's pathogenesis recently underwent a drastic shift. Cutibacterium acnes is a lipophilic bacteria that dominates sebaceous skin sites [3,105]. Each pore is dominated by a single, nearly clonal lineage [105] and individuals have their own unique mix of C. acnes strains [3]. Earlier research suggested that increased sebum production and C. acnes over-proliferation triggered inflammation, abnormal keratinocyte proliferation, and pore duct obstruction [106]. However, healthy individuals have similar to slightly higher C. acnes bioburden [107–110]. Recent works support a dogma where hyper-proliferation of particular C. acnes phylotypes (phylotypes IA1, IA2, 1B1, and IC), reduced C. acnes phylotype diversity, and collective skin microbial dysbiosis triggers the cutaneous inflammation underlying acne development [105,108,111,112]. These acne associated phylotypes tend to induce more inflammation, display elevated porphyrin production [113], and exhibit excessive lipase activity [114]. The fatty acids produced by lipase metabolism of sebum subsequently attract neutrophils and promote hyperkeratosis [114].

Mild acne is generally treated with topical benzoyl peroxide and/or a topical retinoid [115]. Topical then oral antibiotics followed by oral isotretinoin is the mainstay for moderate-severe or refractory acne treatment [115]. Concerns for rising antibiotic resistance and isotretinoin's side effects have driven efforts to identify new therapies designed to prevent or reverse the hyperproliferation of C. acnes strains and equilibrate the microbiome. One promising avenue is selective augmentation of S. epidermidis over C. acnes [116,117] through sucrose supplementation [116] or a probiotic containing encapsulated S. epidermidis and glycerol [118]. Supplementing the skin microbiome with topical probiotics derived from Lactobacillus is another promising avenue [119]. Additional future therapeutics include; i) topical application of anti-microbial peptides [72,120], ii) bacteriophages that strategically infect C. acnes [121], and iii) using oral antibiotics to modulate the gut microbiome and indirectly alter the skin microbiome [122–124].

Homeostasis throughout adulthood

The adult skin microbiome is stable over the span of years [125]. Collectively, established microbial-microbial interaction networks (Table 1), lasting adult skin physiology, and resilient skin immunity maintain balanced adult skin microbial communities. Adult skin microbiomes are dominated by Cutibacterium, Corynebacterium, Staphylococcus and Malassezia species [4]. Each body site has a unique microenvironment and particular sweat and sebaceous gland density that dictates the prevailing microbial composition [4]. Once adulthood is reached, matured and enduring skin physiology promotes consistent sebum production, sweat composition, and surface pH, which collectively provide reliable body site microenvironments and nutrient pools [126,127]. The immune system also reaches full maturation in our early twenties [128]. Unwavering immune function [129,130] further encourages appropriate, reliable responses to our established commensal microbiome and infectious insults. These intrinsic features enable large portions of microbial communities on the skin to persist, despite daily environmental changes [125]. Microbial community stability is evidenced by the longitudinal fixation of highly abundant species (e.g. C. acnes) [125] along with the persistence of several low abundant taxa, which contribute to our unique microbial signature [125,131].

A portion of the taxa within the skin microbiome are also influenced by environmental surroundings. For example, individuals living in Egypt often have a greater abundance of bacteria within the Pseudomonadota phyla [132]; individuals in Cameroon have greater Staphylococcus and Micrococcus [133], those in South Asia tend to have greater abundance of Corynebacterium and Streptococcus [134]; while individuals in Japan have high abundance of Cutibacterium [133]; and Americans and Europeans have greater Corynebacterium species abundance [135]. Collectively this illustrates how individuals living in different geographic locations maintain slightly different skin microbial patterns. Individuals living or working in rural communities also contain more soil and agriculture associated microbes within their skin microbiome [136–138]. We also share more microbial taxa with those living in the same residence [131,139,140] and skin microbiomes become more similar the longer individuals cohabitate [141,142]. In short, the people in our lives, local environment, and geography, partially influence our skin microbial balance. The degree to which a portion of your microbiome shifts with environment changes also appears to be person and situation specific [140–142].

Advancing age

Intriguingly, of the prominent human microbiomes, the skin microbiome is the best predictor of age [143]. With advancing age, distinct skin changes occur including the decline of collagen synthesis, extracellular matrix fragmentation, and a reduction in skin cell regeneration [144]. These changes can manifest as skin wrinkles and more consequently impaired wound healing. Furthermore, these aging related changes shape microbiome composition.

As the skin barrier changes it can lose its ability to retain water, resulting in a compensatory increase in natural moisturizing factor (NFM) production [145]. NMFs both absorb water and can promote bacterial proliferation and adherence to the skin [146]. Subsequently, increased NMFs is associated with greater abundance of numerous taxa, such as Corynebacterium, Micrococcus, Streptococcus, Anaerococcus [127], and a reduction in Cutibacterium [127,144]. Skin microbial diversity also broadly increases [127,137,144,147]. Decreasing sebocyte area and sebum production after menopause in females correlates with loss of Cutibacterium and an increase in Corynebacterium, Streptococcus, Acinetobacter, and Corynebacterium abundance [127,144,148–150]. In males, sebum secretion declines significantly slower so they maintain greater Cutibacterium abundance as they age [137,151].

With age, immune system function also slowly declines [128,130,152,153]. Elderly individuals sustain a low-grade inflammatory state with increased systemic concentrations of pro-inflammatory cytokines [152]. Within the skin Langerhans cells are gradually lost from the epidermis [154]. Cutaneous dendritic cells and the remaining Langerhans also display impaired ability to migrate to lymph nodes and present antigens to T-cells [155,156]. Subsequently, this disrupted antigen presentation along with systemic and local defects in immune signaling ultimately lead to slower immune responses, reduced antimicrobial activity, and impaired wound healing [153,157]. Collectively, impaired immune defenses and increased prevalence of potentially pathogenic bacteria (e.g. beta-hemolytic Streptococci) contribute to the substantial increased risk for skin infection in the elderly and difficulty clearing the infection [153,158,159].

A myriad of dermatologic diseases are associated with advancing age, including dry skin, seborrheic dermatitis, rosacea, disrupted wound healing, and chronic wound infections. Although all these conditions are associated with skin microbiome dysbiosis, specific underlying bacterial and fungal changes remain elusive [18,160,161]. One key exception is seborrheic dermatitis, where recent works find increased abundance of Malassezia and Staphylococcus species to be potential fungal and bacterial biomarkers [162,163]. In tandem with the numerous works seeking to identify avenues that support a balanced, ‘youthful,' microbiome throughout adulthood and later life [144,150,151], greater research into the active role of the microbiome as we age and in the development of age-associated diseases are needed.

There is no single definition of a balanced skin microbiome. While there are core skin microbial members, for any individual, the precise microbial composition is dynamic and unique. This dynamic is influenced by the continuous exchange of microbes with the people we are close to and the world around us. The specific community structure in a particular skin microenvironment is also partially determined by local skin physiology, the microbe–immune interface, and complex microbe–microbe interactions. However, on the human microbiome scale, the largest and somewhat predictable shifts in our skin microbial community compositions occur as we, and our skin, age. Future investigations will continue to elucidate the active role of our dynamic skin microbiome across the lifespan, its implications for dermatologic disease risk and overall health, as well as targeted, microbiome centered therapeutic approaches.

  • The skin and its microbiome are primary interfaces with the environment. The balance of this microbial community influences our risk for various diseases and lifelong health.

  • There is no single definition of a balanced skin microbiome. At a microenvironment level, balance is dictated by the skin niche along with complex host immune-microbe and microbe–microbe interactions. For any individual this balance is also dynamic, shifting with us as we age and navigate the environments around us.

  • Future works in the field of skin microbiology will likely continue to explore i) the active role of our skin microbiome as we age, ii) the complex immune-microbe and microbe–microbe interactions and how they shift over the lifespan, iii) the cutaneous microbiomes of diverse population demographics, iv) the microbiome's influence in protection from and/or disease pathogenesis, and v) the development of novel therapeutic approaches to strategically modulate the microbial balance.

The authors declare that there are no competing interests associated with the manuscript.

This work was supported by grants from the National Institutes of Health (NIAID U19AI142720, NIGMS R35 GM137828 [L.R.K]). The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.

E.C.T and L.R.K contributed to the conceptualization and editing of the text and figures. E. C.T. primarily contributed to the main text and figure creation.

The authors gratefully acknowledge members of the Kalan laboratory for thoughtful feedback on the manuscript. RStudio, Biorender, and Adobe Illustrator were used for figure generation.

3H3MHA

3-hydroxy-3-methylhexanoic acid

3M3SH

3-methyl-3-sullanylhexanol

3SH

3-sulfanylhexanol

BCAA

Branched chain amino acids

CGSC

Cystine-Glycine-S-conjugate

CoNS

Coagulase negative Staphylococcus spp

GC

Glutaminyl-conjugate

1
Costello
,
E.K.
,
Lauber
,
C.L.
,
Hamady
,
M.
,
Fierer
,
N.
,
Gordon
,
J.I.
and
Knight
,
R.
(
2009
)
Bacterial community variation in human body habitats across space and time
.
Science
326
,
1694
1697
2
Grice
,
E.A.
and
Segre
,
J.A.
(
2011
)
The skin microbiome
.
Nat. Rev. Microbiol.
9
,
244
253
3
Oh
,
J.
,
Byrd
,
A.L.
,
Deming
,
C.
,
Conlan
,
S.
,
Kong
,
H.H.
and
Segre
,
J.A.
(
2014
)
Biogeography and individuality shape function in the human skin metagenome
.
Nature
514
,
59
64
4
Byrd
,
A.L.
,
Belkaid
,
Y.
and
Segre
,
J.A.
(
2018
)
The human skin microbiome
.
Nat. Rev. Microbiol.
16
,
143
155
5
Flowers
,
L.
and
Grice
,
E.A.
(
2020
)
The skin microbiota: balancing risk and reward
.
Cell Host Microbe
28
,
190
200
6
Jo
,
J.H.
,
Kennedy
,
E.A.
and
Kong
,
H.H.
(
2017
)
Topographical and physiological differences of the skin mycobiome in health and disease
.
Virulence
8
,
324
333
7
Takahashi
,
T.
and
Gallo
,
R.L.
(
2017
)
The critical and multifunctional roles of antimicrobial peptides in dermatology
.
Dermatol. Clin.
35
,
39
50
8
Uberoi
,
A.
,
Bartow-McKenney
,
C.
,
Zheng
,
Q.
,
Flowers
,
L.
,
Campbell
,
A.
,
Knight
,
S.A.B.
et al (
2021
)
Commensal microbiota regulates skin barrier function and repair via signaling through the aryl hydrocarbon receptor
.
Cell Host Microbe
29
,
1235
1248.e8
9
Dwyer
,
L.R.
and
Scharschmidt
,
T.C.
(
2022
)
Early life host-microbe interactions in skin
.
Cell Host Microbe
30
,
684
695
10
Scharschmidt
,
T.C.
,
Vasquez
,
K.S.
,
Truong
,
H.A.
,
Gearty
,
S.V.
,
Pauli
,
M.L.
,
Nosbaum
,
A.
et al (
2015
)
A wave of regulatory T cells into neonatal skin mediates tolerance to commensal microbes
.
Immunity
43
,
1011
1021
11
Pastar
,
I.
,
O'Neill
,
K.
,
Padula
,
L.
,
Head
,
C.R.
,
Burgess
,
J.L.
,
Chen
,
V.
et al (
2020
)
Staphylococcus epidermidis boosts innate immune response by activation of gamma delta T cells and induction of perforin-2 in human skin
.
Front. Immunol.
11
,
550946
12
Lunjani
,
N.
,
Ahearn-Ford
,
S.
,
Dube
,
F.S.
,
Hlela
,
C.
and
O'Mahony
,
L.
(
2021
)
Mechanisms of microbe-immune system dialogue within the skin
.
Genes Immun.
22
,
276
288
13
Menberu
,
M.A.
,
Liu
,
S.
,
Cooksley
,
C.
,
Hayes
,
A.J.
,
Psaltis
,
A.J.
,
Wormald
,
P.J.
et al (
2021
)
Corynebacterium accolens has antimicrobial activity against Staphylococcus aureus and methicillin-resistant S. aureus pathogens isolated from the sinonasal niche of chronic rhinosinusitis patients
.
Pathogens
10
,
207
14
Claesen
,
J.
,
Spagnolo
,
J.B.
,
Ramos
,
S.F.
,
Kurita
,
K.L.
,
Byrd
,
A.L.
,
Aksenov
,
A.A.
et al (
2020
)
A Cutibacterium acnes antibiotic modulates human skin microbiota composition in hair follicles
.
Sci. Transl. Med.
12
,
eaay5445
15
Chin
,
D.
,
Goncheva
,
M.I.
,
Flannagan
,
R.S.
,
Deecker
,
S.R.
,
Guariglia-Oropeza
,
V.
,
Ensminger
,
A.W.
et al (
2021
)
Coagulase-negative staphylococci release a purine analog that inhibits Staphylococcus aureus virulence
.
Nat. Commun.
12
,
1887
16
O'Sullivan
,
J.N.
,
Rea
,
M.C.
,
O'Connor
,
P.M.
,
Hill
,
C.
and
Ross
,
R.P.
(
2019
)
Human skin microbiota is a rich source of bacteriocin-producing staphylococci that kill human pathogens
.
FEMS Microbiol. Ecol.
95
,
fiy241
17
Sanford
,
J.A.
and
Gallo
,
R.L.
(
2013
)
Functions of the skin microbiota in health and disease
.
Semin. Immunol.
25
,
370
377
18
Verbanic
,
S.
,
Shen
,
Y.
,
Lee
,
J.
,
Deacon
,
J.M.
and
Chen
,
I.A.
(
2020
)
Microbial predictors of healing and short-term effect of debridement on the microbiome of chronic wounds
.
NPJ Biofilms Microbiomes
6
,
21
19
Bjerre
,
R.D.
,
Bandier
,
J.
,
Skov
,
L.
,
Engstrand
,
L.
and
Johansen
,
J.D.
(
2017
)
The role of the skin microbiome in atopic dermatitis: a systematic review
.
Br. J. Dermatol.
177
,
1272
1278
20
Swaney
,
M.H.
and
Kalan
,
L.R.
(
2021
)
Living in your skin: microbes, molecules, and mechanisms
.
Infect. Immun.
89
,
e00695-20
21
Dominguez-Bello
,
M.G.
,
Costello
,
E.K.
,
Contreras
,
M.
,
Magris
,
M.
,
Hidalgo
,
G.
,
Fierer
,
N.
et al (
2010
)
Delivery mode shapes the acquisition and structure of the initial microbiota across multiple body habitats in newborns
.
Proc. Natl Acad. Sci. U.S.A.
107
,
11971
5
22
Chu
,
D.M.
,
Ma
,
J.
,
Prince
,
A.L.
,
Antony
,
K.M.
,
Seferovic
,
M.D.
and
Aagaard
,
K.M.
(
2017
)
Maturation of the infant microbiome community structure and function across multiple body sites and in relation to mode of delivery
.
Nat. Med.
23
,
314
326
23
Younge
,
N.E.
,
Araújo-Pérez
,
F.
,
Brandon
,
D.
and
Seed
,
P.C.
(
2018
)
Early-life skin microbiota in hospitalized preterm and full-term infants
.
Microbiome
6
,
98
24
Ward
,
T.L.
,
Dominguez-Bello
,
M.G.
,
Heisel
,
T.
,
Al-Ghalith
,
G.
,
Knights
,
D.
and
Gale
,
C.A.
(
2018
)
Development of the human mycobiome over the first month of life and across body sites
.
mSystems
3
,
e00140-17
25
Dominguez-Bello
,
M.G.
,
De Jesus-Laboy
,
K.M.
,
Shen
,
N.
,
Cox
,
L.M.
,
Amir
,
A.
,
Gonzalez
,
A.
et al (
2016
)
Partial restoration of the microbiota of cesarean-born infants via vaginal microbial transfer
.
Nat. Med.
22
,
250
253
26
Kennedy
,
E.A.
,
Connolly
,
J.
,
Hourihane
,
J.O.
,
Fallon
,
P.G.
,
McLean
,
W.H.I.
,
Murray
,
D.
et al (
2017
)
Skin microbiome before development of atopic dermatitis: early colonization with commensal staphylococci at 2 months is associated with a lower risk of atopic dermatitis at 1 year
.
J. Allergy Clin. Immunol.
139
,
166
172
27
Casterline
,
B.W.
and
Paller
,
A.S.
(
2021
)
Early development of the skin microbiome: therapeutic opportunities
.
Pediatr. Res.
90
,
731
737
28
King
,
A.
,
Balaji
,
S.
and
Keswani
,
S.G.
(
2013
)
Biology and function of fetal and pediatric skin
.
Facial Plast. Surg. Clin. North Am.
21
,
1
6
29
Visscher
,
M.O.
,
Adam
,
R.
,
Brink
,
S.
and
Odio
,
M.
(
2015
)
Newborn infant skin: physiology, development, and care
.
Clin. Dermatol.
33
,
271
280
30
Manus
,
M.B.
,
Kuthyar
,
S.
,
Perroni-Marañón
,
A.G.
,
Núñez-de la Mora
,
A.
and
Amato
,
K.R.
(
2020
)
Infant skin bacterial communities vary by skin site and infant age across populations in Mexico and the United States
.
mSystems
5
,
e00834-20
31
Schoch
,
J.J.
,
Monir
,
R.L.
,
Satcher
,
K.G.
,
Harris
,
J.
,
Triplett
,
E.
and
Neu
,
J.
(
2019
)
The infantile cutaneous microbiome: a review
.
Pediatr. Dermatol.
36
,
574
580
32
Oranges
,
T.
,
Dini
,
V.
and
Romanelli
,
M.
(
2015
)
Skin physiology of the neonate and infant: clinical implications
.
Adv. Wound Care
4
,
587
595
33
Visscher
,
M.O.
,
Carr
,
A.N.
,
Winget
,
J.
,
Huggins
,
T.
,
Bascom
,
C.C.
,
Isfort
,
R.
et al (
2021
)
Biomarkers of neonatal skin barrier adaptation reveal substantial differences compared to adult skin
.
Pediatr. Res.
89
,
1208
1215
34
Capone
,
K.A.
,
Dowd
,
S.E.
,
Stamatas
,
G.N.
and
Nikolovski
,
J.
(
2011
)
Diversity of the human skin microbiome early in life
.
J. Invest. Dermatol.
131
,
2026
2032
35
Roux
,
P.F.
,
Oddos
,
T.
and
Stamatas
,
G.
(
2022
)
Deciphering the role of skin surface microbiome in skin health: an integrative multiomics approach reveals three distinct metabolite‒microbe clusters
.
J. Invest. Dermatol.
142
,
469
479.e5
36
Oh
,
J.
,
Conlan
,
S.
,
Polley
,
E.C.
,
Segre
,
J.A.
and
Kong
,
H.H.
(
2012
)
Shifts in human skin and nares microbiota of healthy children and adults
.
Genome Med.
4
,
77
37
Jo
,
J.H.
,
Deming
,
C.
,
Kennedy
,
E.A.
,
Conlan
,
S.
,
Polley
,
E.C.
,
Ng
,
W.I.
et al (
2016
)
Diverse human skin fungal communities in children converge in adulthood
.
J. Invest. Dermatol.
136
,
2356
2363
38
Zhu
,
T.
,
Duan
,
Y.Y.
,
Kong
,
F.Q.
,
Galzote
,
C.
and
Quan
,
Z.X.
(
2020
)
Dynamics of skin mycobiome in infants
.
Front. Microbiol.
11
,
1790
39
Paul
,
A.A.
,
Hoffman
,
K.L.
,
Hagan
,
J.L.
,
Sampath
,
V.
,
Petrosino
,
J.F.
and
Pammi
,
M.
(
2020
)
Fungal cutaneous microbiome and host determinants in preterm andterm neonates
.
Pediatr. Res.
88
,
225
233
40
Zhu
,
T.
,
Liu
,
X.
,
Kong
,
F.Q.
,
Duan
,
Y.Y.
,
Yee
,
A.L.
,
Kim
,
M.
et al (
2019
)
Age and mothers: potent influences of children's skin microbiota
.
J. Invest. Dermatol.
139
,
2497
2505.e6
41
Scharschmidt
,
T.C.
,
Vasquez
,
K.S.
,
Pauli
,
M.L.
,
Leitner
,
E.G.
Chu
,
K.
,
Truong
,
H.A.
et al (
2017
)
Commensal microbes and hair follicle morphogenesis coordinately drive treg migration into neonatal skin
.
Cell Host Microbe
21
,
467
477.e5
42
Sanchez Rodriguez
,
R.
,
Pauli
,
M.L.
,
Neuhaus
,
I.M.
,
Yu
,
S.S.
,
Arron
,
S.T.
,
Harris
,
H.W.
et al (
2014
)
Memory regulatory T cells reside in human skin
.
J. Clin. Invest.
124
,
1027
1036
43
Christoph
,
T.
,
Müller-Röver
,
S.
,
Audring
,
H.
,
Tobin
,
D.J.
,
Hermes
,
B.
,
Cotsarelis
,
G.
et al (
2000
)
The human hair follicle immune system: cellular composition and immune privilege
.
Br. J. Dermatol.
142
,
862
873
44
Leech
,
J.M.
,
Dhariwala
,
M.O.
,
Lowe
,
M.M.
,
Chu
,
K.
,
Merana
,
G.R.
,
Cornuot
,
C.
et al (
2019
)
Toxin-triggered interleukin-1 receptor signaling enables early-life discrimination of pathogenic versus commensal skin bacteria
.
Cell Host Microbe
26
,
795
809.e5
45
Gaitanis
,
G.
,
Tsiouri
,
G.
,
Spyridonos
,
P.
,
Stefos
,
Τ
,
Stamatas
,
G.N.
,
Velegraki
,
A.
et al (
2019
)
Variation of cultured skin microbiota in mothers and their infants during the first year postpartum
.
Pediatr. Dermatol.
36
,
460
465
46
Park
,
J.
,
Schwardt
,
N.H.
,
Jo
,
J.H.
,
Zhang
,
Z.
,
Pillai
,
V.
,
Phang
,
S.
et al (
2022
)
Shifts in the skin bacterial and fungal communities of healthy children transitioning through puberty
.
J. Invest. Dermatol.
142
,
212
219
47
Nguyen
,
U.T.
and
Kalan
,
L.R.
(
2022
)
Forgotten fungi: the importance of the skin mycobiome
.
Curr. Opin. Microbiol.
70
,
102235
48
Roslund
,
M.I.
,
Puhakka
,
R.
,
Nurminen
,
N.
,
Oikarinen
,
S.
,
Siter
,
N.
,
Grönroos
,
M.
et al (
2021
)
Long-term biodiversity intervention shapes health-associated commensal microbiota among urban day-care children
.
Environ. Int.
157
,
106811
49
Lehtimäki
,
J.
,
Sinkko
,
H.
,
Hielm-Björkman
,
A.
,
Salmela
,
E.
,
Tiira
,
K.
,
Laatikainen
,
T.
et al (
2018
)
Skin microbiota and allergic symptoms associate with exposure to environmental microbes
.
Proc. Natl Acad. Sci. U.S.A.
115
,
4897
4902
50
Stein
,
M.M.
,
Hrusch
,
C.L.
,
Gozdz
,
J.
,
Igartua
,
C.
,
Pivniouk
,
V.
,
Murray
,
S.E.
et al (
2016
)
Innate immunity and asthma risk in amish and hutterite farm children
.
N. Engl. J. Med.
375
,
411
421
51
Lyons
,
S.A.
,
Knulst
,
A.C.
,
Burney
,
P.G.J.
,
Fernández-Rivas
,
M.
,
Ballmer-Weber
,
B.K.
,
Barreales
,
L.
et al (
2020
)
Predictors of food sensitization in children and adults across Europe
.
J. Allergy Clin. Immunol. Pract.
8
,
3074
3083.e32
52
Meylan
,
P.
,
Lang
,
C.
,
Mermoud
,
S.
,
Johannsen
,
A.
,
Norrenberg
,
S.
,
Hohl
,
D.
et al (
2017
)
Skin colonization by Staphylococcus aureus precedes the clinical diagnosis of atopic dermatitis in infancy
.
J. Invest. Dermatol.
137
,
2497
2504
53
Silverberg
,
J.I.
,
Barbarot
,
S.
,
Gadkari
,
A.
,
Simpson
,
E.L.
,
Weidinger
,
S.
,
Mina-Osorio
,
P.
et al (
2021
)
Atopic dermatitis in the pediatric population: a cross-sectional, international epidemiologic study
.
Ann. Allergy Asthma Immunol.
126
,
417
428.e2
54
Langan
,
S.M.
,
Irvine
,
A.D.
and
Weidinger
,
S.
(
2020
)
Atopic dermatitis
.
Lancet
396
,
345
360
55
Tsakok
,
T.
,
Woolf
,
R.
,
Smith
,
C.H.
,
Weidinger
,
S.
and
Flohr
,
C.
(
2019
)
Atopic dermatitis: the skin barrier and beyond
.
Br. J. Dermatol.
180
,
464
474
56
Seidenari
,
S.
and
Giusti
,
G.
(
1995
)
Objective assessment of the skin of children affected by atopic dermatitis: a study of pH, capacitance and TEWL in eczematous and clinically uninvolved skin
.
Acta Derm. Venereol.
75
,
429
433
57
Jungersted
,
J.M.
,
Scheer
,
H.
,
Mempel
,
M.
,
Baurecht
,
H.
,
Cifuentes
,
L.
,
Høgh
,
J.K.
et al (
2010
)
Stratum corneum lipids, skin barrier function and filaggrin mutations in patients with atopic eczema: filaggrin mutations, skin barrier and lipids
.
Allergy
65
,
911
918
58
Baurecht
,
H.
,
Rühlemann
,
M.C.
,
Rodríguez
,
E.
,
Thielking
,
F.
,
Harder
,
I.
,
Erkens
,
A.S.
et al (
2018
)
Epidermal lipid composition, barrier integrity, and eczematous inflammation are associated with skin microbiome configuration
.
J. Allergy Clin. Immunol.
141
,
1668
1676.e16
59
Berdyshev
,
E.
,
Goleva
,
E.
,
Bronova
,
I.
,
Dyjack
,
N.
,
Rios
,
C.
,
Jung
,
J.
et al (
2018
)
Lipid abnormalities in atopic skin are driven by type 2 cytokines
.
JCI Insight
3
,
e98006
60
Beck
,
L.A.
,
Cork
,
M.J.
,
Amagai
,
M.
,
De Benedetto
,
A.
,
Kabashima
,
K.
,
Hamilton
,
J.D.
et al (
2022
)
Type 2 inflammation contributes to skin barrier dysfunction in atopic dermatitis
.
JID Innov.
2
,
100131
61
Fyhrquist
,
N.
,
Muirhead
,
G.
,
Prast-Nielsen
,
S.
,
Jeanmougin
,
M.
,
Olah
,
P.
,
Skoog
,
T.
et al (
2019
)
Microbe-host interplay in atopic dermatitis and psoriasis
.
Nat. Commun.
10
,
4703
62
Howell
,
M.D.
,
Kim
,
B.E.
,
Gao
,
P.
,
Grant
,
A.V.
,
Boguniewicz
,
M.
,
DeBenedetto
,
A.
et al (
2009
)
Cytokine modulation of atopic dermatitis filaggrin skin expression
.
J. Allergy Clin. Immunol.
124
,
R7
12
63
Hönzke
,
S.
,
Wallmeyer
,
L.
,
Ostrowski
,
A.
,
Radbruch
,
M.
,
Mundhenk
,
L.
,
Schäfer-Korting
,
M.
et al (
2016
)
Influence of Th2 cytokines on the cornified envelope, tight junction proteins, and β-defensins in filaggrin-deficient skin equivalents
.
J. Invest. Dermatol.
136
,
631
639
64
Aziz
,
F.
,
Hisatsune
,
J.
,
Yu
,
L.
,
Kajimura
,
J.
,
Sato'o
,
Y.
,
Ono
,
H.K.
et al (
2020
)
Staphylococcus aureus isolated from skin from atopic-dermatitis patients produces Staphylococcal enterotoxin Y, which predominantly induces T-cell receptor Vα-specific expansion of T cells
.
Infect. Immun.
88
,
e00360-19
65
van Mierlo
,
M.M.F.
,
Pasmans
,
S.G.M.A.
,
Totté
,
J.E.E.
,
de Wit
,
J.
,
Herpers
,
B.L.
,
Vos
,
M.C.
et al (
2021
)
Temporal variation in Staphylococcus aureus protein A genotypes from nose and skin in atopic dermatitis patients
.
Dermatology
237
,
506
512
66
Poh
,
S.E.
,
Koh
,
W.L.C.
,
Lim
,
S.Y.D.
,
Wang
,
E.C.E.
,
Yew
,
Y.W.
,
Common
,
J.E.A.
et al (
2022
)
Expression of Staphylococcus aureus virulence factors in atopic dermatitis
.
JID Innov.
2
,
100130
67
Bjerre
,
R.D.
,
Holm
,
J.B.
,
Palleja
,
A.
,
Sølberg
,
J.
,
Skov
,
L.
and
Johansen
,
J.D.
(
2021
)
Skin dysbiosis in the microbiome in atopic dermatitis is site-specific and involves bacteria, fungus and virus
.
BMC Microbiol.
21
,
256
68
Khadka
,
V.D.
,
Key
,
F.M.
,
Romo-González
,
C.
,
Martínez-Gayosso
,
A.
,
Campos-Cabrera
,
B.L.
,
Gerónimo-Gallegos
,
A.
et al (
2021
)
The skin microbiome of patients with atopic dermatitis normalizes gradually during treatment
.
Front. Cell Infect. Microbiol.
11
,
720674
69
Chng
,
K.R.
,
Tay
,
A.S.L.
,
Li
,
C.
,
Ng
,
A.H.Q.
,
Wang
,
J.
,
Suri
,
B.K.
et al (
2016
)
Whole metagenome profiling reveals skin microbiome-dependent susceptibility to atopic dermatitis flare
.
Nat. Microbiol.
1
,
16106
70
Moosbrugger-Martinz
,
V.
,
Hackl
,
H.
,
Gruber
,
R.
,
Pilecky
,
M.
,
Knabl
,
L.
,
Orth-Höller
,
D.
et al (
2021
)
Initial evidence of distinguishable bacterial and fungal dysbiosis in the skin of patients with atopic dermatitis or netherton syndrome
.
J. Invest. Dermatol.
141
,
114
123
71
Jagielski
,
T.
,
Rup
,
E.
,
Ziółkowska
,
A.
,
Roeske
,
K.
,
Macura
,
A.B.
and
Bielecki
,
J.
(
2014
)
Distribution of Malassezia species on the skin of patients with atopic dermatitis, psoriasis, and healthy volunteers assessed by conventional and molecular identification methods
.
BMC Dermatol.
14
,
3
72
Han
,
S.H.
,
Cheon
,
H.I.
,
Hur
,
M.S.
,
Kim
,
M.J.
,
Jung
,
W.H.
,
Lee
,
Y.W.
et al (
2018
)
Analysis of the skin mycobiome in adult patients with atopic dermatitis
.
Exp. Dermatol.
27
,
366
373
73
Kong
,
H.H.
,
Oh
,
J.
,
Deming
,
C.
,
Conlan
,
S.
,
Grice
,
E.A.
,
Beatson
,
M.A.
et al (
2012
)
Temporal shifts in the skin microbiome associated with disease flares and treatment in children with atopic dermatitis
.
Genome Res.
22
,
850
859
74
Kwon
,
S.
,
Choi
,
J.
,
Shin
,
J.
,
Huh
,
C.
,
Park
,
K.
,
Du
,
M.
et al (
2019
)
Changes in lesional and non-lesional skin microbiome during treatment of atopic dermatitis
.
Acta Derm. Venerol.
99
,
284
290
75
Myles
,
I.A.
,
Williams
,
K.W.
,
Reckhow
,
J.D.
,
Jammeh
,
M.L.
,
Pincus
,
N.B.
,
Sastalla
,
I.
et al (
2016
)
Transplantation of human skin microbiota in models of atopic dermatitis
.
JCI Insight
1
,
e86955
76
Howe,
W.
(
2022
)
Treatment of atopic dermatitis (eczema) [Internet]. UpToDate. [cited 2022 Jun 10]. Available from
: https://www.uptodate.com/contents/treatment-of-atopic-dermatitis-eczema#H1022550
77
Seite
,
S.
,
Flores
,
G.E.
,
Henley
,
J.B.
,
Martin
,
R.
,
Zelenkova
,
H.
,
Aguilar
,
L.
et al (
2014
)
Microbiome of affected and unaffected skin of patients with atopic dermatitis before and after emollient treatment
.
J. Drugs Dermatol.
13
,
1365
1372
PMID:
[PubMed]
78
Liu
,
Y.
,
Wang
,
S.
,
Dai
,
W.
,
Liang
,
Y.
,
Shen
,
C.
,
Li
,
Y.
et al (
2020
)
Distinct skin microbiota imbalance and responses to clinical treatment in children with atopic dermatitis
.
Front. Cell Infect. Microbiol.
10
,
336
79
Olesen
,
C.M.
,
Ingham
,
A.C.
,
Thomsen
,
S.F.
,
Clausen
,
M.L.
,
Andersen
,
P.S.
,
Edslev
,
S.M.
et al (
2021
)
Changes in skin and nasal microbiome and staphylococcal species following treatment of atopic dermatitis with dupilumab
.
Microorganisms
9
,
1487
80
Makrgeorgou
,
A.
,
Leonardi-Bee
,
J.
,
Bath-Hextall
,
F.J.
,
Murrell
,
D.F.
,
Tang
,
M.L.
,
Roberts
,
A.
et al (
2018
)
Probiotics for treating eczema
.
Cochrane Database Syst. Rev.
11
,
CD006135
81
Chang
,
Y.S.
,
Trivedi
,
M.K.
,
Jha
,
A.
,
Lin
,
Y.F.
,
Dimaano
,
L.
and
García-Romero
,
M.T.
(
2016
)
Synbiotics for prevention and treatment of atopic dermatitis: a meta-analysis of randomized clinical trials
.
JAMA Pediatr.
170
,
236
82
Ambrożej
,
D.
,
Kunkiel
,
K.
,
Dumycz
,
K.
and
Feleszko
,
W.
(
2021
)
The use of probiotics and bacteria-derived preparations in topical treatment of atopic dermatitis: a systematic review
.
J. Allergy Clin. Immunol. Pract.
9
,
570
575.e2
83
Myles
,
I.A.
,
Earland
,
N.J.
,
Anderson
,
E.D.
,
Moore
,
I.N.
,
Kieh
,
M.D.
,
Williams
,
K.W.
et al (
2018
)
First-in-human topical microbiome transplantation with Roseomonas mucosa for atopic dermatitis
.
JCI Insight
3
,
120608
84
Nakatsuji
,
T.
,
Hata
,
T.R.
,
Tong
,
Y.
,
Cheng
,
J.Y.
,
Shafiq
,
F.
,
Butcher
,
A.M.
et al (
2021
)
Development of a human skin commensal microbe for bacteriotherapy of atopic dermatitis and use in a phase 1 randomized clinical trial
.
Nat. Med.
27
,
700
709
85
Nakatsuji
,
T.
,
Gallo
,
R.L.
,
Shafiq
,
F.
,
Tong
,
Y.
,
Chun
,
K.
,
Butcher
,
A.M.
et al (
2021
)
Use of autologous bacteriotherapy to treat Staphylococcus aureus in patients with atopic dermatitis: a randomized double-blind clinical trial
.
JAMA Dermatol.
157
,
978
982
86
Williams
,
M.R.
,
Costa
,
S.K.
,
Zaramela
,
L.S.
,
Khalil
,
S.
,
Todd
,
D.A.
,
Winter
,
H.L.
et al (
2019
)
Quorum sensing between bacterial species on the skin protects against epidermal injury in atopic dermatitis
.
Sci. Transl. Med.
11
,
eaat8329
87
Gratton
,
R.
,
Del Vecchio
,
C.
,
Zupin
,
L.
and
Crovella
,
S.
(
2022
)
Unraveling the role of sex hormones on keratinocyte functions in human inflammatory skin diseases
.
Int. J. Mol. Sci.
23
,
3132
88
Marshall
,
W.A.
and
Tanner
,
J.M.
(
1969
)
Variations in pattern of pubertal changes in girls
.
Arch. Dis. Child.
44
,
291
303
89
Marshall
,
W.A.
and
Tanner
,
J.M.
(
1970
)
Variations in the pattern of pubertal changes in boys
.
Arch. Dis. Child.
45
,
13
23
90
Chen
,
F.
,
Hu
,
X.
,
He
,
Y.
and
Huang
,
D.
(
2021
)
Lipidomics demonstrates the association of sex hormones with sebum
.
J. Cosmet. Dermatol.
20
,
2015
2019
91
Strauss
,
J.S.
,
Kligman
,
A.M.
and
Pochi
,
P.E.
(
1962
)
The effect of androgens and estrogens on human sebaceous glands
.
J. Invest. Dermatol.
39
,
139
155
92
Lam
,
T.H.
,
Verzotto
,
D.
,
Brahma
,
P.
,
Ng
,
A.H.Q.
,
Hu
,
P.
,
Schnell
,
D.
et al (
2018
)
Understanding the microbial basis of body odor in pre-pubescent children and teenagers
.
Microbiome
6
,
213
93
Beier
,
K.
,
Ginez
,
I.
and
Schaller
,
H.
(
2005
)
Localization of steroid hormone receptors in the apocrine sweat glands of the human axilla
.
Histochem. Cell Biol.
123
,
61
65
94
Raghunath
,
R.S.
,
Venables
,
Z.C.
and
Millington
,
G.W.M.
(
2015
)
The menstrual cycle and the skin
.
Clin. Exp. Dermatol.
40
,
111
115
95
Wilkinson
,
H.N.
and
Hardman
,
M.J.
(
2017
)
The role of estrogen in cutaneous ageing and repair
.
Maturitas
103
,
60
64
96
Kanda
,
N.
,
Hoashi
,
T.
and
Saeki
,
H.
(
2019
)
The roles of sex hormones in the course of atopic dermatitis
.
Int. J. Mol. Sci.
20
,
4660
97
Yang
,
M.
,
Zhou
,
M.
,
Li
,
Y.
,
Huang
,
H.
and
Jia
,
Y.
(
2021
)
Lipidomic analysis of facial skin surface lipid reveals the causes of pregnancy-related skin barrier weakness
.
Sci. Rep.
11
,
3229
98
James
,
A.G.
,
Austin
,
C.J.
,
Cox
,
D.S.
,
Taylor
,
D.
and
Calvert
,
R.
(
2013
)
Microbiological and biochemical origins of human axillary odour
.
FEMS Microbiol. Ecol.
83
,
527
540
99
Troccaz
,
M.
,
Gaïa
,
N.
,
Beccucci
,
S.
,
Schrenzel
,
J.
,
Cayeux
,
I.
,
Starkenmann
,
C.
et al (
2015
)
Mapping axillary microbiota responsible for body odours using a culture-independent approach
.
Microbiome
3
,
3
100
Emter
,
R.
and
Natsch
,
A.
(
2008
)
The sequential action of a dipeptidase and a β-lyase is required for the release of the human body odorant 3-methyl-3-sulfanylhexan-1-ol from a secreted cys-gly-(S) conjugate by corynebacteria
.
J. Biol. Chem.
283
,
20645
20652
101
Baumann
,
T.
,
Bergmann
,
S.
,
Schmidt-Rose
,
T.
,
Max
,
H.
,
Martin
,
A.
,
Enthaler
,
B.
et al (
2014
)
Glutathione-conjugated sulfanylalkanols are substrates for ABCC 11 and γ -glutamyl transferase 1: a potential new pathway for the formation of odorant precursors in the apocrine sweat gland
.
Exp. Dermatol.
23
,
247
252
102
Chen
,
L.
,
Li
,
J.
,
Zhu
,
W.
,
Kuang
,
Y.
,
Liu
,
T.
,
Zhang
,
W.
et al (
2020
)
Skin and gut microbiome in psoriasis: gaining insight into the pathophysiology of it and finding novel therapeutic strategies
.
Front. Microbiol.
11
,
589726
103
Wark
,
K.J.L.
and
Cains
,
G.D.
(
2021
)
The microbiome in hidradenitis suppurativa: a review
.
Dermatol. Ther. (Heidelb)
11
,
39
52
104
Heng
,
A.H.S.
and
Chew
,
F.T.
(
2020
)
Systematic review of the epidemiology of acne vulgaris
.
Sci. Rep.
10
,
5754
105
Conwill
,
A.
,
Kuan
,
A.C.
,
Damerla
,
R.
,
Poret
,
A.J.
,
Baker
,
J.S.
,
Tripp
,
A.D.
et al (
2022
)
Anatomy promotes neutral coexistence of strains in the human skin microbiome
.
Cell Host Microbe
30
,
171
182.e7
106
Dawson
,
A.L.
and
Dellavalle
,
R.P.
(
2013
)
Acne vulgaris
.
BMJ
346
,
f2634
107
Barnard
,
E.
,
Shi
,
B.
,
Kang
,
D.
,
Craft
,
N.
and
Li
,
H.
(
2016
)
The balance of metagenomic elements shapes the skin microbiome in acne and health
.
Sci. Rep.
6
,
39491
108
Fitz-Gibbon
,
S.
,
Tomida
,
S.
,
Chiu
,
B.H.
,
Nguyen
,
L.
,
Du
,
C.
,
Liu
,
M.
et al (
2013
)
Propionibacterium acnes strain populations in the human skin microbiome associated with acne
.
J. Invest. Dermatol.
133
,
2152
2160
109
Dessinioti
,
C.
and
Katsambas
,
A.
(
2017
)
Propionibacterium acnes and antimicrobial resistance in acne
.
Clin. Dermatol.
35
,
163
167
110
Pécastaings
,
S.
,
Roques
,
C.
,
Nocera
,
T.
,
Peraud
,
C.
,
Mengeaud
,
V.
,
Khammari
,
A.
et al (
2018
)
Characterisation of Cutibacterium acnes phylotypes in acne and in vivo exploratory evaluation of Myrtacine®
.
J. Eur. Acad. Dermatol. Venereol.
32
,
15
23
111
Dagnelie
,
M.
,
Montassier
,
E.
,
Khammari
,
A.
,
Mounier
,
C.
,
Corvec
,
S.
and
Dréno
,
B.
(
2019
)
Inflammatory skin is associated with changes in the skin microbiota composition on the back of severe acne patients
.
Exp. Dermatol.
28
,
961
967
112
Dagnelie
,
M.A.
,
Corvec
,
S.
,
Saint-Jean
,
M.
,
Nguyen
,
J.M.
,
Khammari
,
A.
and
Dréno
,
B.
(
2019
)
Cutibacterium acnes phylotypes diversity loss: a trigger for skin inflammatory process
.
J. Eur. Acad. Dermatol. Venereol.
33
,
2340
2348
113
Barnard
,
E.
,
Johnson
,
T.
,
Ngo
,
T.
,
Arora
,
U.
,
Leuterio
,
G.
,
McDowell
,
A.
et al (
2020
)
Porphyrin production and regulation in cutaneous propionibacteria
.
mSphere
5
,
e00793-19
114
Sanford
,
J.A.
,
O'Neill
,
A.M.
,
Zouboulis
,
C.C.
and
Gallo
,
R.L.
(
2019
)
Short-chain fatty acids from Cutibacterium acnes activate both a canonical and epigenetic inflammatory response in human sebocytes
.
J. Immunol.
202
,
1767
1776
115
Zaenglein
,
A.L.
(
2018
)
Acne vulgaris
.
N. Engl. J. Med.
379
,
1343
1352
116
Wang
,
Y.
,
Kao
,
M.S.
,
Yu
,
J.
,
Huang
,
S.
,
Marito
,
S.
,
Gallo
,
R.
et al (
2016
)
A precision microbiome approach using sucrose for selective augmentation of staphylococcus epidermidis fermentation against Propionibacterium acnes
.
Int. J. Mol. Sci.
17
,
1870
117
Marito
,
S.
,
Keshari
,
S.
,
Traisaeng
,
S.
,
My
,
D.T.T.
,
Balasubramaniam
,
A.
,
Adi
,
P.
et al (
2021
)
Electricity-producing Staphylococcus epidermidis counteracts Cutibacterium acnes
.
Sci. Rep.
11
,
12001
118
Yang
,
A.J.
,
Marito
,
S.
,
Yang
,
J.J.
,
Keshari
,
S.
,
Chew
,
C.H.
,
Chen
,
C.C.
et al (
2018
)
A microtube array membrane (MTAM) encapsulated live fermenting Staphylococcus epidermidis as a skin probiotic patch against Cutibacterium acnes
.
Int. J. Mol. Sci.
20
,
14
119
Sathikulpakdee
,
S.
,
Kanokrungsee
,
S.
,
Vitheejongjaroen
,
P.
,
Kamanamool
,
N.
,
Udompataikul
,
M.
and
Taweechotipatr
,
M.
(
2022
)
Efficacy of probiotic-derived lotion from Lactobacillus paracasei MSMC 39-1 in mild to moderate acne vulgaris, randomized controlled trial
.
J. Cosmet. Dermatol.
21
,
5092
5097
120
Woodburn
,
K.W.
,
Jaynes
,
J.
and
Clemens
,
L.E.
(
2020
)
Designed antimicrobial peptides for topical treatment of antibiotic resistant acne vulgaris
.
Antibiotics
9
,
23
121
Castillo
,
D.E.
,
Nanda
,
S.
and
Keri
,
J.E.
(
2019
)
Propionibacterium (Cutibacterium) acnes bacteriophage therapy in acne: current evidence and future perspectives
.
Dermatol. Ther. (Heidelb)
9
,
19
31
122
Jung
,
G.W.
,
Tse
,
J.E.
,
Guiha
,
I.
and
Rao
,
J.
(
2013
)
Prospective, randomized, open-label trial comparing the safety, efficacy, and tolerability of an acne treatment regimen with and without a probiotic supplement and minocycline in subjects with mild to moderate acne
.
J. Cutan. Med. Surg.
17
,
114
122
123
Fabbrocini
,
G.
,
Bertona
,
M.
,
Picazo
,
Ó.
,
Pareja-Galeano
,
H.
,
Monfrecola
,
G.
and
Emanuele
,
E.
(
2016
)
Supplementation with Lactobacillus rhamnosus SP1 normalises skin expression of genes implicated in insulin signalling and improves adult acne
.
Benef. Microbes
7
,
625
630
124
Thompson
,
K.G.
,
Rainer
,
B.M.
,
Antonescu
,
C.
,
Florea
,
L.
,
Mongodin
,
E.F.
,
Kang
,
S.
et al (
2020
)
Minocycline and its impact on microbial dysbiosis in the skin and gastrointestinal tract of acne patients
.
Ann. Dermatol.
32
,
21
30
125
Oh
,
J.
,
Byrd
,
A.L.
,
Park
,
M.
,
Kong
,
H.H.
and
Segre
,
J.A.
(
2016
)
Temporal stability of the human skin microbiome
.
Cell
165
,
854
866
126
Luebberding
,
S.
,
Krueger
,
N.
and
Kerscher
,
M.
(
2013
)
Skin physiology in men and women: in vivo evaluation of 300 people including TEWL, SC hydration, sebum content and skin surface pH
.
Int. J. Cosmet. Sci.
35
,
477
483
127
Howard
,
B.
,
Bascom
,
C.C.
,
Hu
,
P.
,
Binder
,
R.L.
,
Fadayel
,
G.
,
Huggins
,
T.G.
et al (
2022
)
Aging-associated changes in the adult human skin microbiome and the host factors that affect skin microbiome composition
.
J. Invest. Dermatol.
142
,
1934
1946.e21
128
Simon
,
A.K.
,
Hollander
,
G.A.
and
McMichael
,
A.
(
2015
)
Evolution of the immune system in humans from infancy to old age
.
Proc. R. Soc. B.
282
,
20143085
129
Weitzmann
,
A.
,
Naumann
,
R.
,
Dudeck
,
A.
,
Zerjatke
,
T.
,
Gerbaulet
,
A.
and
Roers
,
A.
(
2020
)
Mast cells occupy stable clonal territories in adult steady-state skin
.
J. Invest. Dermatol.
140
,
2433
2441.e5
130
Tokura
,
Y.
,
Phadungsaksawasdi
,
P.
,
Kurihara
,
K.
,
Fujiyama
,
T.
and
Honda
,
T.
(
2021
)
Pathophysiology of skin resident memory T cells
.
Front. Immunol.
11
,
618897
131
Richardson
,
M.
,
Gottel
,
N.
,
Gilbert
,
J.A.
and
Lax
,
S.
(
2019
)
Microbial similarity between students in a common dormitory environment reveals the forensic potential of individual microbial signatures
.
mBio
10
,
e01054-19
132
Ramadan
,
M.
,
Solyman
,
S.
,
Taha
,
M.
and
Hanora
,
A.
(
2016
)
Preliminary characterization of human skin microbiome in healthy Egyptian individuals
.
Cell. Mol. Biol.
62
,
21
27
PMID:
[PubMed]
133
Ogai
,
K.
,
Nana
,
B.C.
,
Lloyd
,
Y.M.
,
Arios
,
J.P.
,
Jiyarom
,
B.
,
Awanakam
,
H.
et al (
2022
)
Skin microbiome profile of healthy Cameroonians and Japanese
.
Sci. Rep.
12
,
1364
134
Chaudhari
,
D.S.
,
Dhotre
,
D.P.
,
Agarwal
,
D.M.
,
Gaike
,
A.H.
,
Bhalerao
,
D.
,
Jadhav
,
P.
et al (
2020
)
Gut, oral and skin microbiome of Indian patrilineal families reveal perceptible association with age
.
Sci. Rep.
10
,
5685
135
Cho
,
H.W.
and
Eom
,
Y.B.
(
2021
)
Forensic analysis of human microbiome in skin and body fluids based on geographic location
.
Front. Cell Infect. Microbiol.
11
,
695191
136
Peng
,
M.
and
Biswas
,
D.
(
2020
)
Environmental influences of high-density agricultural animal operation on human forearm skin microflora
.
Microorganisms
8
,
E1481
137
Ying
,
S.
,
Zeng
,
D.N.
,
Chi
,
L.
,
Tan
,
Y.
,
Galzote
,
C.
,
Cardona
,
C.
et al (
2015
)
The influence of age and gender on skin-associated microbial communities in urban and rural human populations
.
PLoS ONE
10
,
e0141842
138
Hospodsky
,
D.
,
Pickering
,
A.J.
,
Julian
,
T.R.
,
Miller
,
D.
,
Gorthala
,
S.
,
Boehm
,
A.B.
et al (
2014
)
Hand bacterial communities vary across two different human populations
.
Microbiology
160
,
1144
1152
139
Lax
,
S.
,
Smith
,
D.P.
,
Hampton-Marcell
,
J.
,
Owens
,
S.M.
,
Handley
,
K.M.
,
Scott
,
N.M.
et al (
2014
)
Longitudinal analysis of microbial interaction between humans and the indoor environment
.
Science
345
,
1048
1052
140
Leung
,
M.H.Y.
,
Tong
,
X.
,
Wilkins
,
D.
,
Cheung
,
H.H.L.
and
Lee
,
P.K.H.
(
2018
)
Individual and household attributes influence the dynamics of the personal skin microbiota and its association network
.
Microbiome
6
,
26
141
Sharma
,
A.
,
Richardson
,
M.
,
Cralle
,
L.
,
Stamper
,
C.E.
,
Maestre
,
J.P.
,
Stearns-Yoder
,
K.A.
et al (
2019
)
Longitudinal homogenization of the microbiome between both occupants and the built environment in a cohort of United States Air Force Cadets
.
Microbiome
7
,
70
142
Voorhies
,
A.A.
,
Mark Ott
,
C.
,
Mehta
,
S.
,
Pierson
,
D.L.
,
Crucian
,
B.E.
,
Feiveson
,
A.
et al (
2019
)
Study of the impact of long-duration space missions at the International Space Station on the astronaut microbiome
.
Sci. Rep.
9
,
9911
143
Huang
,
S.
,
Haiminen
,
N.
,
Carrieri
,
A.P.
,
Hu
,
R.
,
Jiang
,
L.
,
Parida
,
L.
et al (
2020
)
Human skin, oral, and gut microbiomes predict chronological age
.
mSystems
5
,
e00630-19
144
Kim
,
H.J.
,
Kim
,
J.J.
,
Myeong
,
N.R.
,
Kim
,
T.
,
Kim
,
D.
,
An
,
S.
et al (
2019
)
Segregation of age-related skin microbiome characteristics by functionality
.
Sci. Rep.
9
,
16748
145
Boireau-Adamezyk
,
E.
,
Baillet-Guffroy
,
A.
and
Stamatas
,
G.N.
(
2021
)
The stratum corneum water content and natural moisturization factor composition evolve with age and depend on body site
.
Int. J. Dermatol.
60
,
834
839
146
Feuillie
,
C.
,
Vitry
,
P.
,
McAleer
,
M.A.
,
Kezic
,
S.
,
Irvine
,
A.D.
,
Geoghegan
,
J.A.
et al (
2018
)
Adhesion of Staphylococcus aureus to corneocytes from atopic dermatitis patients is controlled by natural moisturizing factor levels
.
mBio
9
,
e01184-18
147
Dimitriu
,
P.A.
,
Iker
,
B.
,
Malik
,
K.
,
Leung
,
H.
,
Mohn
,
W.W.
and
Hillebrand
,
G.G.
(
2019
)
New insights into the intrinsic and extrinsic factors that shape the human skin microbiome
.
mBio
10
,
e00839-19
148
Shibagaki
,
N.
,
Suda
,
W.
,
Clavaud
,
C.
,
Bastien
,
P.
,
Takayasu
,
L.
,
Iioka
,
E.
et al (
2017
)
Aging-related changes in the diversity of women's skin microbiomes associated with oral bacteria
.
Sci. Rep.
7
,
10567
149
Jugé
,
R.
,
Rouaud-Tinguely
,
P.
,
Breugnot
,
J.
,
Servaes
,
K.
,
Grimaldi
,
C.
,
Roth
,
M.P.
et al (
2018
)
Shift in skin microbiota of Western European women across aging
.
J. Appl. Microbiol.
125
,
907
916
150
Kim
,
H.J.
,
Oh
,
H.N.
,
Park
,
T.
,
Kim
,
H.
,
Lee
,
H.G.
,
An
,
S.
et al (
2022
)
Aged related human skin microbiome and mycobiome in Korean women
.
Sci. Rep.
12
,
2351
151
Alkema
,
W.
,
Boekhorst
,
J.
,
Eijlander
,
R.T.
,
Schnittger
,
S.
,
De Gruyter
,
F.
,
Lukovac
,
S.
et al (
2021
)
Charting host-microbe co-metabolism in skin aging and application to metagenomics data
.
PLoS ONE
16
,
e0258960
152
Brodin
,
P.
and
Davis
,
M.M.
(
2017
)
Human immune system variation
.
Nat. Rev. Immunol.
17
,
21
29
153
Chambers
,
E.S.
and
Vukmanovic-Stejic
,
M.
(
2020
)
Skin barrier immunity and ageing
.
Immunology
160
,
116
125
154
Hasegawa
,
T.
,
Feng
,
Z.
,
Yan
,
Z.
,
Ngo
,
K.H.
,
Hosoi
,
J.
and
Demehri
,
S.
(
2020
)
Reduction in human epidermal langerhans cells with age is associated with decline in CXCL14-mediated recruitment of CD14+ monocytes
.
J. Invest. Dermatol.
140
,
1327
1334
155
Pilkington
,
S.M.
,
Ogden
,
S.
,
Eaton
,
L.H.
,
Dearman
,
R.J.
,
Kimber
,
I.
and
Griffiths
,
C.E.M.
(
2018
)
Lower levels of interleukin-1 β gene expression are associated with impaired Langerhans’ cell migration in aged human skin
.
Immunology
153
,
60
70
156
Grolleau-Julius
,
A.
,
Harning
,
E.K.
,
Abernathy
,
L.M.
and
Yung
,
R.L.
(
2008
)
Impaired dendritic cell function in aging leads to defective antitumor immunity
.
Cancer Res.
68
,
6341
6349
157
Choi
,
E.H.
(
2019
)
Aging of the skin barrier
.
Clin. Dermatol.
37
,
336
345
158
Cranendonk
,
D.R.
,
Lavrijsen
,
A.P.M.
,
Prins
,
J.M.
and
Wiersinga
,
W.J.
(
2017
)
Cellulitis: current insights into pathophysiology and clinical management
.
Neth. J. Med.
75
,
366
378
PMID:
[PubMed]
159
Haran
,
J.P.
,
Wilsterman
,
E.
,
Zeoli
,
T.
,
Beaudoin
,
F.L.
,
Tjia
,
J.
and
Hibberd
,
P.L.
(
2017
)
Elderly patients are at increased risk for treatment failure in outpatient management of purulent skin infections
.
Am. J. Emerg. Med.
35
,
249
254
160
Daou
,
H.
,
Paradiso
,
M.
,
Hennessy
,
K.
and
Seminario-Vidal
,
L.
(
2021
)
Rosacea and the microbiome: a systematic review
.
Dermatol. Ther. (Heidelb)
11
,
1
12
161
Liu
,
C.
,
Ponsero
,
A.J.
,
Armstrong
,
D.G.
,
Lipsky
,
B.A.
and
Hurwitz
,
B.L.
(
2020
)
The dynamic wound microbiome
.
BMC Med.
18
,
358
162
Lin
,
Q.
,
Panchamukhi
,
A.
,
Li
,
P.
,
Shan
,
W.
,
Zhou
,
H.
,
Hou
,
L.
et al (
2021
)
Malassezia and Staphylococcus dominate scalp microbiome for seborrheic dermatitis
.
Bioprocess Biosyst. Eng.
44
,
965
975
163
Tao
,
R.
,
Li
,
R.
and
Wang
,
R.
(
2021
)
Skin microbiome alterations in seborrheic dermatitis and dandruff: a systematic review
.
Exp. Dermatol.
30
,
1546
1553
164
Fujimura
,
S.
and
Nakamura
,
T.
(
1978
)
Purification and properties of a bacteriocin-like substance (Acnecin) of oral Propionibacterium acnes
.
Antimicrob. Agents Chemother.
14
,
893
898
165
Wang
,
Y.
,
Dai
,
A.
,
Huang
,
S.
,
Kuo
,
S.
,
Shu
,
M.
,
Tapia
,
C.P.
et al (
2014
)
Propionic acid and its esterified derivative suppress the growth of methicillin-resistant Staphylococcus aureus USA300
.
Benef. Microbes
5
,
161
168
166
Shu
,
M.
,
Wang
,
Y.
,
Yu
,
J.
,
Kuo
,
S.
,
Coda
,
A.
,
Jiang
,
Y.
et al (
2013
)
Fermentation of Propionibacterium acnes, a commensal bacterium in the human skin microbiome, as skin probiotics against methicillin-resistant Staphylococcus aureus
.
PLoS ONE
8
,
e55380
167
Gribbon
,
E.M.
,
Cunliffe
,
W.J.
and
Holland
,
K.T.
(
1993
)
Interaction of Propionibacterium acnes with skin lipids in vitro
.
J. Gen. Microbiol.
139
,
1745
1751
168
Brüggemann
,
H.
,
Salar-Vidal
,
L.
,
Gollnick
,
H.P.M.
and
Lood
,
R.
(
2021
)
A Janus-faced bacterium: host-beneficial and -detrimental roles of Cutibacterium acnes
.
Front. Microbiol.
12
,
673845
169
Megyeri
,
K.
,
Orosz
,
L.
,
Bolla
,
S.
,
Erdei
,
L.
,
Rázga
,
Z.
,
Seprényi
,
G.
et al (
2018
)
Propionibacterium acnes induces autophagy in keratinocytes: involvement of multiple mechanisms
.
J. Invest. Dermatol.
138
,
750
759
170
Kao
,
H.J.
,
Wang
,
Y.H.
,
Keshari
,
S.
,
Yang
,
J.J.
,
Simbolon
,
S.
,
Chen
,
C.C.
et al (
2021
)
Propionic acid produced by Cutibacterium acnes fermentation ameliorates ultraviolet B-induced melanin synthesis
.
Sci. Rep.
11
,
11980
171
Schwarz
,
A.
,
Bruhs
,
A.
and
Schwarz
,
T.
(
2017
)
The short-chain fatty acid sodium butyrate functions as a regulator of the skin immune system
.
J. Invest. Dermatol.
137
,
855
864
172
Sanford
,
J.A.
,
Zhang
,
L.J.
,
Williams
,
M.R.
,
Gangoiti
,
J.A.
,
Huang
,
C.M.
and
Gallo
,
R.L.
(
2016
)
Inhibition of HDAC8 and HDAC9 by microbial short-chain fatty acids breaks immune tolerance of the epidermis to TLR ligands
.
Sci. Immunol.
1
,
eaah4609
173
Keshari
,
S.
,
Balasubramaniam
,
A.
,
Myagmardoloonjin
,
B.
,
Herr
,
D.R.
,
Negari
,
I.P.
and
Huang
,
C.M.
(
2019
)
Butyric acid from probiotic staphylococcus epidermidis in the skin microbiome down-regulates the ultraviolet-induced pro-inflammatory IL-6 cytokine via short-chain fatty acid receptor
.
Int. J. Mol. Sci.
20
,
4477
174
Horn
,
K.J.
,
Jaberi Vivar
,
A.C.
,
Arenas
,
V.
,
Andani
,
S.
,
Janoff
,
E.N.
and
Clark
,
S.E.
(
2022
)
Corynebacterium species inhibit Streptococcus pneumoniae colonization and infection of the mouse airway
.
Front. Microbiol.
12
,
804935
175
Bomar
,
L.
,
Brugger
,
S.D.
,
Yost
,
B.H.
,
Davies
,
S.S.
and
Lemon
,
K.P.
(
2016
)
Corynebacterium accolens releases antipneumococcal free fatty acids from human nostril and skin surface triacylglycerols
.
mBio
7
,
e01725-15
176
Ramsey
,
M.M.
,
Freire
,
M.O.
,
Gabrilska
,
R.A.
,
Rumbaugh
,
K.P.
and
Lemon
,
K.P.
(
2016
)
Staphylococcus aureus shifts toward commensalism in response to corynebacterium species
.
Front. Microbiol.
7
,
1230
177
Hardy
,
B.L.
,
Dickey
,
S.W.
,
Plaut
,
R.D.
,
Riggins
,
D.P.
,
Stibitz
,
S.
,
Otto
,
M.
et al (
2019
)
Corynebacterium pseudodiphtheriticum exploits Staphylococcus aureus virulence components in a novel polymicrobial defense strategy
.
mBio
10
,
e02491-18
178
Lynch
,
D.
,
O'Connor
,
P.M.
,
Cotter
,
P.D.
,
Hill
,
C.
,
Field
,
D.
and
Begley
,
M.
(
2019
)
Identification and characterisation of capidermicin, a novel bacteriocin produced by Staphylococcus capitis
.
PLoS ONE
14
,
e0223541
179
Kumar
,
R.
,
Jangir
,
P.K.
,
Das
,
J.
,
Taneja
,
B.
and
Sharma
,
R.
(
2017
)
Genome analysis of Staphylococcus capitis TE8 reveals repertoire of antimicrobial peptides and adaptation strategies for growth on human skin
.
Sci. Rep.
7
,
10447
180
O'Neill
,
A.M.
,
Nakatsuji
,
T.
,
Hayachi
,
A.
,
Williams
,
M.R.
,
Mills
,
R.H.
,
Gonzalez
,
D.J.
et al (
2020
)
Identification of a human skin commensal bacterium that selectively kills Cutibacterium acnes
.
J. Invest. Dermatol.
140
,
1619
1628.e2
181
Paharik
,
A.E.
,
Parlet
,
C.P.
,
Chung
,
N.
,
Todd
,
D.A.
,
Rodriguez
,
E.I.
,
Van Dyke
,
M.J.
et al (
2017
)
Coagulase-negative staphylococcal strain prevents staphylococcus aureus colonization and skin infection by blocking quorum sensing
.
Cell Host Microbe
22
,
746
756.e5
182
Iwase
,
T.
,
Uehara
,
Y.
,
Shinji
,
H.
,
Tajima
,
A.
,
Seo
,
H.
,
Takada
,
K.
et al (
2010
)
Staphylococcus epidermidis Esp inhibits Staphylococcus aureus biofilm formation and nasal colonization
.
Nature
465
,
346
349
183
Nakatsuji
,
T.
,
Chen
,
T.H.
,
Narala
,
S.
,
Chun
,
K.A.
,
Two
,
A.M.
,
Yun
,
T.
et al (
2017
)
Antimicrobials from human skin commensal bacteria protect against Staphylococcus aureus and are deficient in atopic dermatitis
.
Sci. Transl. Med.
9
,
eaah4680
184
Hernández-Aristizábal
,
I.
and
Ocampo-Ibáñez
,
I.D.
(
2021
)
Antimicrobial peptides with antibacterial activity against vancomycin-resistant Staphylococcus aureus strains: classification, structures, and mechanisms of action
.
Int. J. Mol. Sci.
22
,
7927
185
Zipperer
,
A.
,
Konnerth
,
M.C.
,
Laux
,
C.
,
Berscheid
,
A.
,
Janek
,
D.
,
Weidenmaier
,
C.
et al (
2016
)
Human commensals producing a novel antibiotic impair pathogen colonization
.
Nature
535
,
511
516
186
Bitschar
,
K.
,
Sauer
,
B.
,
Focken
,
J.
,
Dehmer
,
H.
,
Moos
,
S.
,
Konnerth
,
M.
et al (
2019
)
Lugdunin amplifies innate immune responses in the skin in synergy with host- and microbiota-derived factors
.
Nat. Commun.
10
,
2730
187
Brown
,
M.M.
,
Kwiecinski
,
J.M.
,
Cruz
,
L.M.
,
Shahbandi
,
A.
,
Todd
,
D.A.
,
Cech
,
N.B.
et al (
2020
)
Novel peptide from commensal Staphylococcus simulans blocks methicillin-resistant Staphylococcus aureus quorum sensing and protects host skin from damage
.
Antimicrob. Agents Chemother.
64
,
e00172-20
188
Severn
,
M.
,
Cho
,
Y.S.K.
,
Manzer
,
H.
,
Bunch
,
Z.
,
Shahbandi
,
A.
,
Todd
,
D.
et al (
2022
)
The commensal Staphylococcus warneri makes peptide inhibitors of MRSA quorum sensing that protect skin from atopic or necrotic damage
.
J. Invest. Dermatol.
142
,
3349
3352.e5
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).