You are on page 1of 8

The Association of the Skin Microbiota With

Health, Immunity, and Disease


M Egert1, R Simmering2 and CU Riedel3

The human skin is densely colonized by a highly diverse microbiota comprising all three domains of life. Long believed to
represent mainly a source of infection, the human skin microbiota is nowadays well accepted as an important driver of
human (skin) health and well-being. This microbiota is influenced by many host and environmental factors and interacts
closely with the skin immune system. Although cause and effect are usually difficult to discriminate, changes in the skin
microbiota clearly play a role in the pathobiology of many types of skin disease and cosmetic disorders. Consequently, treat-
ment and prevention strategies have to respect this role, rendering pre- and probiotic and even transplantation therapies an
additional option to the use of antibiotics.

THE HUMAN SKIN: NOT ONE BUT MANY MICROBIAL Structural and functional characteristics of the human skin
HABITATS The human skin provides a mechanical and biological barrier
The human skin represents one of the largest and most versatile against chemical, physical, and pathogenic threats, thereby main-
organs of the human body. It functions as a protective interface taining host homeostasis. It participates in thermoregulation, sup-
between the largely sterile interior of the human body and the ports immunological functions, and protects against UV
unsterile outer environment. Like other epithelial interfaces with radiation by melanogenesis. Anatomically, the skin comprises
the external environment, the human skin is densely colonized three distinct compartments (Figure 1): the stratum corneum,
with a complex microbial community.1 The entire microbial consisting of dead, keratinized epithelial cells, the avascular epi-
community of a given habitat is also referred to as “microbiota” dermis, mainly composed of (living) keratinocytes, and the der-
or “microbiome.” While “microbiota” rather refers to the micro- mis, a fibroblast-rich network of collagen and elastin fibers
bial taxa associated with a given environment, a “microbiome” providing strength and elasticity. The dermis also contains capil-
comprises the catalog of these microbes and their genetic material lary and lymphatic vessels, which serve as the entry and exit por-
(DNA/RNA). However, both terms are often used interchange- tals for immune cells. Additional skin appendages such as hair
ably.2 For the sake of consistency, and because we largely focus follicles, sebaceous glands, and sweat glands, as well as nerve end-
ings, are also found in the dermis.4
on the interactions of different microbial taxa with the human
Several factors contribute to the protective function of the skin
host, we use “microbiota” throughout this text. “Microflora” is
against colonization by pathogenic microorganisms and, in turn,
another widely used term for “microbial community,” particularly have to be overcome by the resident, commensal, or even mutual-
in medical circles. However, it originates from a time when istic members of the skin microbiota. The relatively low tempera-
microbiology was a subdiscipline of botany, and microorganisms ture of the skin (29–348C compared to 378C in the core of the
were categorized as a kind of plant. Therefore, we suggest avoid- human body) and its acidic pH of about 4.5 to 5.5 represent
ing this term. rather unfavorable conditions for most skin pathogenic bacteria.
For a long time, this skin microbiota was seen as a source of Moreover, the continuous proliferation of epithelial cells in the
contamination and infection. However, the human microbiome dermis and shedding of dead, keratinized cells in the epidermis
project and concomitant molecular, i.e., cultivation- provide protection against microbial infection and degradation.4
independent, research projects strongly suggested that the In addition to these chemical and physical mechanisms, the bar-
human skin microbiota is of major importance for human health rier function of the skin is supported by a wealth of different cell
and well-being.1,3 types that provide both innate and adaptive immunity.5
1
Faculty of Medical and Life Sciences, Institute of Precision Medicine, Microbiology and Hygiene Group, Furtwangen University, Villingen-Schwenningen,
Germany. 2Henkel AG & Co. KGaA, Corporate Scientific Services, Du €sseldorf, Germany. 3Institute of Microbiology and Biotechnology, University of Ulm, Ulm,
Germany. Correspondence: M Egert (Markus.Egert@hs-furtwangen.de)
Received 20 February 2017; accepted 28 March 2017; advance online publication 5 April 2017. doi:10.1002/cpt.698

62 VOLUME 102 NUMBER 1 | JULY 2017 | www.cpt-journal.com


Figure 1 Layers of the human skin and associated glands and vessels. Contrary to the historical notion that microbial life is restricted to the epidermis
and its appendages (hair follicles and glands), recent evidence suggest (skin) microorganisms to be also present in deeper layers. Interaction between
microbes and Langerhans cells leads to the induction of FoxP31 regulatory T cells (Treg). In deeper layers of the skin, contact of microorganisms with con-
ventional dendritic cell (cDC) subsets induces production of IL-1, IL-17, and IFN-g by T cells, migration of these T cells to the epidermis, activation of natu-
ral killer (NK) cells, and secretion of antimicrobial peptides (AMPs) by keratinocytes.

Secretion of sweat and sebum eccrine, apocrine, and intermediate, i.e., apoeccrine sweat glands
In particular, the uneven distribution of different gland types cre- (Figure 1).
ates many different habitats across the human body, which differ Eccrine sweat glands are the most abundant sweat glands of
significantly in environmental conditions. Sebaceous parts, such the human body. On average, 100 to 200 glands per cm2 are dis-
as the scalp, forehead, neck, and the upper part of the back, are tributed virtually all over the body surface. While palms and soles
greasy and contain a large number of sebaceous glands responsible show higher densities (600 glands per cm2), other parts of the
for secretion of sebum. A few hours after birth, secretion of body, e.g., lips and nail beds, are depleted of eccrine glands.8
sebum strongly increases, which lasts for a couple of days and Thermoregulatory perspiration by eccrine sweat glands is effective
then decreases again. Normally, in late childhood (starting from birth on and is affected by environmental parameters, such
approximately at the age of 9 years), sebum production rises again as temperature, humidity, skin and body temperature, but also by
and lasts until the adult level is reached. The sebaceous gland is a physical fitness, circadian rhythm, and the menstrual cycle. Secre-
target organ, but also an important production site of hormones, tion by eccrine glands is mainly triggered by temperature but also
especially of active androgens.6 by pain, stress, fear, and anxiety, resulting in emotional sweating.
While some parts of the human body, such as armpits, the Digestion is also speculated to induce eccrine sweating. Another
genital area, and feet, are rather occluded, moist, and warm, crucial function of eccrine sweat is the control of microbial colo-
there are also large areas on forearms, legs, and the lower part of nization and growth by acidification of the skin surface.1 Eccrine
the back which are relatively exposed and dry. Perspiration, i.e., sweat is a largely translucent liquid mainly composed of water,
secretion through sweat glands, is crucial for the regulation of sodium and potassium salts, as well as amino acids, sugars, lactate,
body temperature, which represents an important factor for and glycoproteins.9 Its exact composition depends on hormonal
body homeostasis.7,8 Sweat glands are specialized exocrine glands activity, physical condition, acclimatization to environmental
that are appendages of the skin and can be categorized into conditions, as well as secretion rate.8

CLINICAL PHARMACOLOGY & THERAPEUTICS | VOLUME 102 NUMBER 1 | JULY 2017 63


Apoeccrine sweat glands are far less abundant on skin than unknown mechanisms.11 However, their effects are by far not
eccrine glands. About 2,000 of them are distributed around the restricted to killing of microorganisms, but also include immuno-
eyes and ears and on chest skin, while the highest densities are modulatory functions. Some AMPs interact directly with Toll-
reported for axillae and groin. Apocrine glands release their secre- like receptors (TLRs), a family of innate recognition receptors,
tion into hair canals, and are restricted to hairy body surfaces. and this interaction seems to limit excessive inflammatory
They are activated with puberty and androgen stimulation and responses.11 Moreover, LL-37 and hBDs act as potent chemotac-
are then stimulated by hormones.7–9 tic signals and attract monocytes, dendritic cells, neutrophils, and
Apocrine sweat appears milky and viscous, and is known to T cells.11
comprise lipids, lactates, nitrogen, electrolytes, steroids, proteins, Similar to the gastrointestinal immune system, the skin is rich
and vitamins in addition to other ions.8 However, the exact com- in a wide range of different cell populations of the innate and
position of apocrine secretion has yet to be elucidated due to the adaptive immune system even in the absence of a pathogenic or
lack of pure samples. Apocrine glands are suspected to be inflammatory stimulus (Figure 1). Cells of the adaptive immune
involved in emotional sweating, but their original function system in the skin include antigen-presenting cells (APCs),
remains unidentified.7 In contrast to eccrine sweat glands, apo- CD41, and CD81 T cells.24 Several populations of APCs are
crine glands secret substances that can be transformed into mal- present in different layers of the skin, of which epidermal Langer-
odorous molecules upon lysis by bacteria and thus contribute hans cells are the most abundant.25 These different populations
significantly to development of human body odor.7–10 of APCs have different functional properties. Langerhans cells
seem to have a rather immunosuppressive phenotype that coun-
The immune system of the skin terbalances excessive T-cell responses induced by conventional
The skin harbors a wide range of different cell populations dendritic cell (cDC) populations of the dermis.25 At the other
including keratinocytes, macrophages, dendritic cells, innate lym- end of the spectrum, cDC subsets prime na€ıve T cells to differen-
phocytes, and different T-cell populations that belong to or per- tiate into appropriate effector T cells, depending on the stimu-
form functions of the innate and adaptive immunity. The lus.25 As observed in the gastrointestinal tract, the skin
effector mechanisms of these cells include production of antimi- microbiota plays a major role in maintaining the skin immune
crobial compounds, expression of innate recognition receptors to system in homeostasis under normal conditions. The skin, but
sense and respond to microbe-associated molecular patterns, and not the gut microbiota, is required to induce appropriate levels of
more specialized, antigen-specific responses.5 The primary innate interleukin (IL) 17 and interferon g-producing T cells and
defense system of the human skin consists of an arsenal of anti- FoxP3-expressing regulatory T-cell populations.26 Early exposure
microbial peptides (AMPs), and enzymes mainly produced by of human neonates to S. epidermidis induces S. epidermidis-
keratinocytes.11 Enzymes with antimicrobial activity found on specific, FoxP31 regulatory T cells in the skin which are thought
human skin include lysozyme, RNAses, and S100 family pro- to contribute to inhibition of excessive (proinflammatory)
teins.11 The main AMPs of the human skin are human b- immune responses to this commensal skin bacterium.27 More-
defensins (hBD-1, hBD-2, hBD-3, hBD-4), and the cathelicidin over, different dermal DC subsets cooperate in antigen presenta-
LL-37.11 b-defensins are small cationic peptides (usually 4– tion and induction of IL-17-producing CD8-positive T cells in
5 kDa) with a characteristic set of three disulfide bonds.12 LL-37 response to S. epidermidis in an IL-1-dependent manner. These
is a member of the cathelicidin family, which are host defense IL-17-producing T cells then migrate to the epidermis and stimu-
peptides of vertebrates, of which LL-37 is the only human repre- late expression of AMPs by keratinocytes.28
sentative. The active form of LL37 is derived from its precursor
hCAP18 by proteolytic cleavage, consists of 37 amino acids with THE SKIN MICROBIOTA IN HEALTH AND DISEASE
two leucine residues at the N-terminus, and has a broad range Structure of the skin microbiota of healthy adults
antimicrobial activity against bacteria, fungi, and viruses.13 In the last decade, high-throughput sequencing technologies have
Expression of AMPs by keratinocytes is regulated by various greatly facilitated the in-depth taxonomic analysis of the commu-
mechanisms. While hBD-1 is generally considered as being con- nity structure of the human skin microbiota. It is now well estab-
stitutively expressed in the epithelium, hBD-2, hBD-3, and lished that this microbiota comprises prokaryotes, eukaryotes
LL-37 are induced during inflammation, e.g., in acne lesions and (fungi, metazoic parasites), and viruses.1,3,29,30 The prokaryotic
psoriasis,14–16 and in response to microbial stimuli.17–20 Activa- community is dominated by Bacteria, however, Archaea are
tion occurs via cytokines of the interleukin-1 family and the clearly also present.31 Thus, the skin microbiota comprises all
nuclear factor kappa B (NF-jB) pathway.21,22 Polymorphisms in three domains of life. Metazoic parasites (mostly arthropods, e.g.,
an NF-jB binding site in the promoter of hBD-1 are associated mites) and viruses will not be discussed here, as they do not repre-
with reduced expression of hBD-1 and hBD-3 and increased sent microorganisms in a narrower sense (mites) or living micro-
nasal carriage of S. aureus.23 Expression of LL-37 is induced by organisms (viruses), respectively.
vitamin D3, which is produced upon exposure to UVB light, in Assuming a skin surface of 2 m2 and maximum microbial
combination with inflammatory stimuli and skin injury.11 (bacterial) cell densities of 106 cm-2, the skin of healthy adults
The primary role of antimicrobial peptides and proteins of the can be estimated to be colonized by a maximum of 1010 micro-
skin is to inactivate microorganisms by disruption of membrane organisms in total. However, bacteria are distributed very
integrity, enzymatic degradation of the cell wall, or other, yet unevenly across the different niches of the human skin. Cell

64 VOLUME 102 NUMBER 1 | JULY 2017 | www.cpt-journal.com


numbers range from 102 cm22 (fingertips, back) to 106 cm22 detected; however, Actinobacteria, Firmicutes, and Proteobacteria
(forehead, axilla). Owing to its physicochemical conditions, the again accounted for 94% of the sequences; the most abundant
human skin is typically colonized by mesophilic, xerophilic, aci- genera were Propionibacterium (31.6% of all sequences), Strepto-
dophilic, osmotolerant, and facultative aerobic microorganisms. coccus (17.2%), Staphylococcus (8.3%), Corynebacterium (4.3%),
However, depending on the respective niche, also microbes with and Lactobacillus (3.1%). Hands from the same individual shared
other physiological traits can occur. While it was long believed only 17% of their phylotypes, with different individuals sharing
that on healthy skin microbial life is restricted to epidermis, hair only 13%. Women had a significantly higher diversity than men,
follicles, and sebaceous and sweat glands, recent analyses32 suggest and community composition was significantly affected by hand-
that viable microbes are found also in deeper skin layers, i.e., the edness, time since last hand washing, and an individual’s gender.
dermis and underlying fat tissue (Figure 1). This finding is While most studies on the skin microbiota addressed the epi-
important from an immunological point of view, because it sug- dermis and its appendages, knowledge about the microbiota of
gests direct communication between host and microbial cells in a subepidermal compartments of normal skin is scarce. Using a
tissue previously thought to be sterile, unless injured and contam- multiphasic approach,32 probably active bacteria have been
inated with bacteria from upper skin parts or the environment. reported from the dermis and dermal adipose tissue. Interestingly,
At present, molecular high-throughput studies (see below, and this microbiota was dominated by Proteobacteria, while Actino-
Supplementary Table S1) allowed detection of more than 25 bacteria and Firmicutes were much less abundant. Although fur-
bacterial phyla on human skin, although mostly in low abun- ther studies are needed to substantiate and better interpret this
dance. The vast majority of skin bacteria are affiliated with three finding, theses data suggest that the commensal skin microbiota
phyla: Actinobacteria, Firmicutes, and Proteobacteria. Moreover, extends within the dermis, which allows physical contact between
the human skin microbiota proved to be individual (particularly bacteria and various cells below the basement membrane, i.e.,
on genus and species-level), body site-dependent, and to change that normal commensal bacterial communities directly communi-
with time and age. cate with the host in a tissue previously thought to be sterile.
Investigated over a period of 8–10 months, the superficial volar Finally, still relatively little is known about nonbacterial mem-
forearms of six healthy adults contained 247 operational taxo- bers of the human skin microbiota, e.g., fungi and archaea. Of 14
nomic units (OTUs, here used in the sense of “species”) belong- investigated skin sites in 10 healthy adults, 11 core-body and arm
ing to 119 genera and 10 phyla.33 On average, 48 species were sites were dominated by Malassezia yeasts, showing differences
detected per human individual. Actinobacteria, Firmicutes, and only on the species level. By contrast, three foot sites showed a
Proteobacteria accounted for 95% of the clones. Only four higher fungal diversity, comprising also genera such as Aspergillus,
(3.4%) of the 119 genera (Propionibacterium, Corynebacterium, Cryptococcus, Rhodotorula, Epicoccum, and others, and a lower sta-
Staphylococcus, Streptococcus) were observed in those subjects, who bility over time. Fungal communities were also influenced
were tested twice over the period, albeit these genera represented by skin topography, but—in comparison to bacterial skin
54% of all clones. communities—less clearly by skin physiology.36 For a long time,
Over a period of 4–6 months, 20 different body sites of 10 Archaea were thought to be absent from human skin. However, a
healthy adults harbored 19 bacterial phyla, but most sequences recent report31 suggested that they can represent more than 4% of
were assigned to just four of them: Actinobacteria (52%), Firmi- the human prokaryotic skin microbiota. Phylotypes detected by
cutes, (24%), Proteobacteria (17%), and Bacteroidetes. (6%).34 Of polymerase chain reaction (PCR) and fluorescence in situ hybridi-
205 identified genera, represented by at least five sequences, three zation (FISH) were mostly affiliated with Thaumarchaeota, and
accounted for more than 62% of the sequences: corynebacteria to a lesser extent also with Euryarchaeota. So far, their physiologi-
(23%, Actinobacteria), propionibacteria (23%; Actinobacteria), cal role is obscure. According to the function of Thaumarchaeota
and staphylococci (17%; Firmicutes). Sebaceous sites were domi- in many environmental ecosystems, a role in skin ammonia
nated by propionibacteria and staphylococci. Corynebacteria and metabolism has been suggested.36
staphylococci were predominant in moist habitats. Rather unex- A wealth of intrinsic (host) and extrinsic (environmental,
pectedly, in particular in view of earlier cultivation-dependent or behavioral) factors have been identified that influence the com-
small-scale cloning-based community analyses, a mixed popula- position of the skin microbiota of the human body on a spatial
tion of bacteria resided in dry sites, with a greater prevalence of and temporal scale.1,3,37 Intrinsic factors include age, gender,
b-Proteobacteria and Flavobacteriales. Species diversity and tem- genetic predisposition, immune status, underlying diseases (skin
poral stability of the skin microbiota were also site-dependent: and nonskin disease), skin site, and microbe–microbe interac-
the latter was higher for protected sites (nares, external auditory tions. Extrinsic factors include, for instance, life style, domestic
canal) than for exposed sites (buttock, popliteal). The effect of and personal hygiene, cohabitation, and environmental factors
skin niche on the composition of the skin microbiota was more such as place of living (geography) and solar radiation. Supple-
dependent on the investigated skin site than on the individual. mentary Table S1 represents an updated version of a recently
The left and right palmar surfaces of 27 healthy men and 24 published table37 containing select examples of recent studies
healthy women contained more than 4,700 bacterial phylotypes investigating various factors that shape the overall composition of
(“species”) representing an overall microbial diversity similar to the skin microbiota. In addition to the factors mentioned in
that of the human intestinal tract.35 On average, every hand was Table S1, several others are likely to influence the composition
colonized by 158 different phylotypes. More than 25 phyla were of the human skin microbiota, such as diet, climate, occupation

CLINICAL PHARMACOLOGY & THERAPEUTICS | VOLUME 102 NUMBER 1 | JULY 2017 65


(clothing), stress, etc.; however, appropriate studies are still responses elicited by injured skin cells.45 So far, skin bacteria were
missing. believed to cause rather than to reduce skin inflammation. Also,
S. epidermidis was shown to enhance skin defense mechanisms
Protective functions of the human skin microbiota against infection by enhancing gene expression of antimicrobial
Given the high interindividual variation in the composition of peptides, such as hBDs.46 Finally, it was shown that S. epidermidis
the (skin) microbiotas of healthy subjects, it is difficult to define can tune the function of resident skin T-lymphocytes and
a “healthy” skin. However, it is generally believed that a healthy thereby contribute to protective immunity against skin patho-
skin microbiota is characterized by a considerable diversity of gens. Monoassociation of the skin of previously germ-free mice
commensal or even beneficial (symbiotic) bacteria.38 In contrast, with S. epidermidis restored the production of IL-17A in skin T-
illness is thought be associated with a disturbed (imbalanced) cells and thereby protective immunity against infection.26 Clearly,
microbiota characterized by a loss of microbial diversity and skin commensals such as S. epidermidis are important drivers and
increased absolute numbers or relative abundance of pathogenic amplifiers of human skin immunity44 and a good example that
bacteria; a state called dysbiosis.39 Colonization of the skin with
the human body comprises additional niches to the intestinal
a balanced or healthy microbiota is generally recognized as benefi-
tract, were immunosurveillance systems are locally fine-tuned by
cial because it is thought to contribute to protection against
a commensal microbiota.47
infections, a phenomenon known as colonization resistance.
Moreover, the microbiota may also play a role in other, noninfec- The skin microbiota in skin disorders and diseases
tious disease, e.g., by immunomodulatory effects. However, given Many human skin disorders and diseases have been linked with
the pronounced temporal, spatial, and interindividual variations changes in skin microbiota composition and/or functionality.
in microbial diversity, assigning beneficial functions to distinct However, in most cases it is still not clear whether these changes
taxonomic groups of the complex skin microbiota, which also are the cause or the consequence of the underlying disease. Never-
comprises large numbers of yet uncultured species, is challenging. theless, such changes are expected to have diagnostic, preventive,
The acidic pH of normal healthy skin, also termed the “acid and potentially therapeutic implications.
mantle,” represents an important line of defense against patho- In the case of acne vulgaris, P. acnes is regarded as the primary
genic microorganisms. Facultative anaerobes, such as Propionibac- disease-associated bacterium. Importantly, the major trigger of
terium acnes, reside in sebaceous glands where their lipolytic
acne is not the presence of P. acnes itself, but hormone-induced
enzymes release free fatty acids from sebum. Together with
increased sebum production, providing P. acnes with optimal liv-
sebum, these free fatty acids are continuously shed onto the skin
ing conditions in an anaerobic and lipid-rich environment.
surface and contribute to the acidic pH.1
Genome analyses of P. acnes revealed a variety of virulence factors
Another mechanism by which commensal bacteria contribute
probably involved in the pathobiology of acne, e.g., hyaluroni-
to colonization resistance is the production of bacteriocins and
dases, lipases, and proteases.48 In a recent study49 comparing the
other antimicrobial factors, which are produced to increase com-
skin of acne patients with healthy individuals, it was shown that
petitive fitness in a densely populated environment. With respect
the relative abundances of P. acnes species were similar, but on
to the skin ecosystem, staphylococci, and in particular S. epider-
midis, can be regarded as the best-investigated bacteriocin- the strain-level the population structures were significantly differ-
producers.40 Some commensal strains of. S. epidermidis can ent in the two cohorts. Certain strains were highly associated
protect their host from colonization with S. aureus by means of a with acne, and other strains were enriched in healthy skin. The
serine protease (Esp) that destroys S. aureus biofilms; for instance, authors also identified potential genetic determinants of various
in the anterior nares.41 Nasal S. lugdunensis was recently shown to P. acnes strains in association with acne or health, which could
produce lugdunin, a cyclic peptide antibiotic that inhibits nasal become future targets for therapeutic interventions. For instance,
colonization with S. aureus.42 Lugdunin represents a bactericidal particularly disease-associated strains contained sequences encod-
activity acting against various important pathogens, including ing CRISPRs (Clustered Regularly Interspaced Short Palin-
methicillin-resistant S. aureus, vancomycin-resistant Enterococcus dromic Repeats), which have been shown to confer protection
faecalis, Listeria monocytogenes, Streptococcus pneumoniae, and against viruses, phages, and plasmids.
Pseudomonas aeruginosa. Although peptide bacteriocins such as Higher frequencies of Firmicutes and lower frequencies of Acti-
lugdunin are rather ineffective alternatives to systemic antibiotics nobacteria, in particular of propionibacteria, were detected in pso-
due to their immunogenicity, they may represent potential antibi- riatic lesions compared to normal skin stretches of patients and
otic agents for topical applications. In this case, either the pure of skin from healthy individuals.50 An analysis of triplet samples
bacteriocin or a probiotic strain producing the compound could from diseased, unaffected, and healthy skin showed that psoriasis
be applied to the skin. induced physiological changes both at the lesion site and the sys-
Similar to the human gut, the skin microbiota is thought to be temic level.51 Lesions were characterized by a reduced microbial
involved in the development and regulation of the innate and species diversity. Psoriasis patients were characterized by higher
adaptive immune system and maintenance of skin homeosta- combined relative abundances of the major skin genera Coryne-
sis.43,44 Lipoteichoic acids of S. epidermidis prevented skin injury- bacterium, Propionibacterium, Staphylococcus, and Streptococcus,
based inflammation by inhibition of inflammatory cytokines while genera such as Cupriavidus, Methylobacterium, and Schlege-
release from keratinocytes as well as TLR2-based immune- lella were significantly less abundant.

66 VOLUME 102 NUMBER 1 | JULY 2017 | www.cpt-journal.com


Atopic dermatitis (AD) is characterized by increased bacterial
colonization with S. aureus.52,53 Interestingly, the induction of
hBD-2 and hBD-3 usually seen in inflammatory lesions of acne
and psoriasis is not observed in AD lesion.15,16,54 Since both pep-
tides show potent antimicrobial activity against S. aureus,17 a lack
of appropriate induction of hBD2 and hBD3 expression might
contribute to increased levels of S. aureus, and thus pathology of
AD. At present, the mechanistic defects leading to insufficient
induction of hBD-2 and hBD-3 have not been elucidated.
Rosacea and seborrheic dermatitis are skin diseases linked with
Demodex mites and Malassezia fungi and an immunological and
microbial dysbiosis of the skin ecosystem.3 In the case of rosacea,
microbiota-associated changes on the skin and in the small intes-
tine have been observed simultaneously.55 Seborrheic dermatitis,
particularly of the scalp (dandruff), affects 50% of the global
population and is mainly caused by M. restricta and M. globosa
producing high levels of irritant fatty acids, such as oleic acid, Figure 2 The three major concepts to manipulate the human skin micro-
causing skin hyperproliferation and scaling. In unaffected per- biota in order to improve human health and well-being. The use of topically
sons, these organisms are harmless commensals of the skin and applied antibiotics is effective in reducing the overall bacterial load of the
scalp microbiota. The factors that allow these normally harmless skin, but does not allow for a differentiation between beneficial (green)
commensals to switch to a pathogenic state are still not fully and adverse/pathogenic (red) bacteria. In contrast, pre- and probiotic
strategies aim at increasing the number of beneficial bacteria and reduc-
understood. Recent evidence shows the increased production of ing the number of adverse/pathogenic ones. Skin microbiota transplanta-
specific phospholipases on affected skin sites in dandruff and sig- tion is a very new strategy and is based on the transfer of beneficial
naling molecules such as malassezin in seborrheic dermatitis.56 microorganisms or communities from healthy to diseased and microbiolog-
In addition to dandruff, also other, rather cosmetic, skin prob- ically imbalanced skin. All picture elements were obtained from pixabay.
lems such as impure skin, sensitive skin, or body odor are linked com and are distributed under a Creative Commons CC0 license. The cen-
tral skin element was changed from a picture created by Don Bliss,
with the skin microbiota. Strong body odor (bromidrosis) is a National Cancer Institute, Rockville, MD, USA.
pathological phenomenon, triggers a person’s attractiveness for
mosquitoes, and might thus play a role in the transmission of uncouple mutualistic host–microbe relationships. Recently, orally
infectious diseases such as malaria.57 The human armpit repre-
administered vancomycin was shown to influence the amount
sents one of the most densely colonized habitats of the human
and composition of bacteria on the skin of mice, with negative
skin. Its composition is gender-specific58 and characterized by an
consequences for wound healing.65 Thus, alternative therapies
asymmetric (left vs. right) activity.59 Members of the resident
that make use of the mutualistic interactions between the skin
microbiota metabolize odorless sweat to malodorous compounds
microbiota and its host are increasingly investigated.64,66–68
using hydrolytic enzyme activities, such as aminoacylase and cys-
The prebiotic concept was transferred from nutrition to the
tathionin-b-lyase.10 While it was long believed that mainly cory-
field of skin care products about 15 years ago.69,70 For example,
nebacteria produce body odor, recent cultivation-independent
application of a cosmetic product containing 0.5% of a mixture
studies58 suggest that also other groups are involved, such as anae-
of pine, black current, and ginseng to human skin twice per day
rococci60 and staphylococci.61,62
for 3 weeks was effective in inhibiting the growth of P. acnes,
MANIPULATION OF THE HUMAN SKIN MICROBIOTA whereas commensals such as S. epidermidis were not affected.70
Strategies to manipulate the skin microbiota for therapeutic rea- Farnesol and xylitol were used to balance the skin microbiota of
sons can be divided into three categories (Figure 2): antimicro- patients with atopic dermatitis, by removing and preventing the
bial (antibiotic) therapy, application of pre- or probiotics, or adhesion of biofilm-producing S. aureus species.71 An extract
transplantation of entire microbial consortia to diseased skin. from avocado, consisting mainly of mannoheptulose and persei-
While antibiotic treatment aims at an elimination of specific tol, affected the adherence of M. furfur to keratinocytes, and
(pathogenic) microorganisms, the goal of the application of pro- induced the production of hBD2,72 possibly due to structural
biotics or microbiota transplantation is an increase of selected, similarities of its main components to cell wall constituents of
potentially beneficial microorganisms.63,64 yeasts. Future developments in this field have to balance a stimu-
The overall beneficial effect of antibiotics against severe micro- lated antimicrobial defense system vs. an immunological overreac-
bial skin infections is unquestionable. While their aim is to tion. This issue was addressed in a study on the induction of
reduce or completely eradicate specific pathogens, they often also defensins in keratinocytes by plant extracts.73 Only some of the
reduce or eliminate other, sometimes beneficial, members of the tested extracts, e.g., arnica, betel, black elder, and mugwort,
microbiota. Moreover, these effects, and specifically the use of enhanced expression of defensins without inducing proinflamma-
broad-spectrum antibiotics, may have pervasive and long-lasting tory cytokines, such as IL-8, IL-1a, or MIP-3a, rendering them
consequences for the microbial ecosystem of the skin that potential cosmetic or skin therapeutic ingredients.

CLINICAL PHARMACOLOGY & THERAPEUTICS | VOLUME 102 NUMBER 1 | JULY 2017 67


The topical application of beneficial bacteria as probiotics (as CONFLICT OF INTEREST
living or inactivated whole cells) represents a second strategy to R.S. is employed by Henkel AG & Co, KGaA, a company which produces
(re)balance the skin microbiota. In an early study,74 the use of skin cosmetics. M.E. has previously worked for Henkel and is still
involved in microbiology projects funded by the company.
food-grade propionibacteria was proposed for cosmetic products,
since cutaneous isolates might be correlated with skin infections.
C 2017 American Society for Clinical Pharmacology and Therapeutics
V
The selected strains revealed an antimicrobial activity against sev-
eral skin pathogens, such as M. furfur, C. albicans, and S. aureus,
attributed to the secretion of organic acids and/or competition 1. Grice, E.A. & Segre, J.A. The skin microbiome. Nat. Rev. Microbiol. 9,
244–253 (2011).
for attachment sites on keratin. 2. Ursell, L.K., Metcalf, J.L., Parfrey, L.W. & Knight, R. Defining the
In a pilot study on keratoconjunctivitis, a reduction of symp- human microbiome. Nutr. Rev. 70, S38–44 (2012).
toms and of the molecular markers ICAM-1 and TLR-4, both 3. Schommer, N.N. & Gallo, R.L. Structure and function of the human
skin microbiome. Trends Microbiol. 21, 660–668 (2013).
proteins of the innate immune system, were observed after admin- 4. Wilson, M. Bacteriology of Humans an Ecological Perspective
istration of eyedrops containing inactivated Lactobacillus acidophi- (Blackwell, Malden, MA; 2008).
lus cells for 2–4 weeks.75 Besides lactic acid bacteria, biomass of 5. Heath, W.R. & Carbone, F.R. The skin-resident and migratory immune
system in steady state and memory: innate lymphocytes, dendritic
Vitreoscilla filiformis represents another probiotic ingredient in cos- cells and T cells. Nat. Immunol. 14, 978–985 (2013).
metic products. It was shown to modulate a mitochondrial super- 6. Zouboulis, C.C. Die Talgdru €se. Hautarzt 61, 467–477 (2010).
oxide dismutase in skin cells, thereby conferring antioxidative 7. Wilke, K., Martin, A., Terstegen, L. & Biel, S.S. A short history of
sweat gland biology. Int. J. Cosmet. Sci. 29, 169–179 (2007).
protection.76 Biomass of the same organism lead to a stimulation 8. Noe €l, F., Pie
rard-Franchimont, C., Pierard, G.E. & Quatresooz, P.
of mRNA expression coding for antimicrobial peptides hBD2 and Sweaty skin, background and assessments. Int. J. Dermatol. 51,
S100A7 in a reconstructed epidermis model. Furthermore, involve- 647–655 (2012).
9. Kelly, D.P. & Wood, A.P. Skin microbiology, body odor, and
ment of the TLR2/protein kinase C zeta-transduction pathway methylotrophic bacteria. In Handbook of Hydrocarbon and Lipid
was supposed.77 Finally, the effects of orally ingested probiotics on Microbiology (ed. Timmis, K.N.) 3203–3213 (Springer, Heidelberg;
the skin were described: In a double-blinded, placebo-controlled 2010) (Biomedical and Life Sciences).
10. Fredrich, E., Barzantny, H., Brune, I. & Tauch, A. Daily battle against
trial, ingestion of Lactococcus lactis strains improved selected skin body odor: towards the activity of the axillary microbiota. Trends
and body properties in women, such skin elasticity.78 Microbiol. 21, 305–312 (2013).
In contrast to the intestinal tract, where fecal transplantation 11. Gallo, R.L. & Hooper, L.V. Epithelial antimicrobial defence of the skin
and intestine. Nat. Rev. Immunol. 12, 503–516 (2012).
has become an increasingly accepted therapeutic option, the 12. Mattar, E.H., Almehdar, H.A., Yacoub, H.A., Uversky, V.N. & Redwan,
transplantation of (selected members of) a healthy skin micro- E.M. Antimicrobial potentials and structural disorder of human and
biota to diseased skin tissue with an unbalanced microbiota is still animal defensins. Cytokine Growth Factor Rev. 28, 95–111 (2016).
13. Xhindoli, D. et al. The human cathelicidin LL-37—A pore-forming
at a very experimental stage. Very recently, cultivable Gram- antibacterial peptide and host-cell modulator. Biochim. Biophys. Acta.
negative bacteria taken from healthy volunteers but not from 1858, 546–566 (2016).
patients with AD were shown to be associated with enhanced 14. Frohm, M. et al. The expression of the gene coding for the
antibacterial peptide LL-37 is induced in human keratinocytes during
barrier function, innate immunity activation, and control of S. inflammatory disorders. J. Biol. Chem. 272, 15258–15263 (1997).
aureus in in vitro and in vivo models of AD.63 15. Nomura, I. et al. Cytokine milieu of atopic dermatitis, as compared to
psoriasis, skin prevents induction of innate immune response genes.
OUTLOOK: TRENDS AND CHALLENGES J. Immunol. 171, 3262–3269 (2003).
16. Kelha € la
€, H.-L. et al. IL-17/Th17 pathway is activated in acne lesions.
The microbiota of the human skin clearly plays a significant role PLoS One 9, e105238 (2014).
for human health and well-being. Future research in this field will 17. Midorikawa, K. et al. Staphylococcus aureus susceptibility to innate
therefore (continue) to focus on the definition of a “normal” antimicrobial peptides, beta-defensins and CAP18, expressed by
human keratinocytes. Infect. Immun. 71, 3730–3739 (2003).
(healthy) state of skin microbiota composition and functioning and 18. Nagy, I. et al. Propionibacterium acnes and lipopolysaccharide induce
its shaping by environmental and host factors. The interplay of the the expression of antimicrobial peptides and proinflammatory
three domains of life and of viruses among each other and with cytokines/chemokines in human sebocytes. Microbes Infect. 8,
2195–2205 (2006).
host physiology and immune system is another important topic, 19. Percoco, G. et al. Antimicrobial peptides and pro-inflammatory
not only on the human skin, but also regarding interactions with cytokines are differentially regulated across epidermal layers
the microbiota of other body compartments, e.g., the human intes- following bacterial stimuli. Exp. Dermatol. 22, 800–806 (2013).
20. Feng, Z. et al. Epithelial innate immune response to Acinetobacter
tinal tract. In order to better discriminate cause and effect, a deeper baumannii challenge. Infect. Immun. 82, 4458–4465 (2014).
and more mechanistic (functional) understanding of the role of the 21. Johnston, A. et al. IL-1F5, -F6, -F8, and -F9: a novel IL-1 family
skin microbiota in skin diseases and cosmetic skin disorders is signaling system that is active in psoriasis and promotes
keratinocyte antimicrobial peptide expression. J. Immunol. 186,
needed. Increased understanding in this area will aid the definition 2613–2622 (2011).
of microbiological biomarkers for an early and reliable diagnosis of 22. Towne, J.E., Garka, K.E., Renshaw, B.R., Virca, G.D. & Sims, J.E.
skin diseases as well as the development of alternative therapeutic Interleukin (IL)-1F6, IL-1F8, and IL-1F9 signal through IL-1Rrp2 and IL-
1RAcP to activate the pathway leading to NF-kappaB and MAPKs. J.
strategies (prebiotic, probiotic, transplantation) and products which Biol. Chem. 279, 13677–13688 (2004).
respect and utilize mutualistic host–microbe relationships. 23. Nurjadi, D., Herrmann, E., Hinderberger, I. & Zanger, P. Impaired b-
defensin expression in human skin links DEFB1 promoter
Additional Supporting Information may be found in the online version of polymorphisms with persistent Staphylococcus aureus nasal
this article. carriage. J. Infect. Dis. 207, 666–674 (2013).

68 VOLUME 102 NUMBER 1 | JULY 2017 | www.cpt-journal.com


24. Belkaid, Y. & Tamoutounour, S. The influence of skin microorganisms 54. Nomura, I. et al. Distinct patterns of gene expression in the skin
on cutaneous immunity. Nat. Rev. Immunol. 16, 353–366 (2016). lesions of atopic dermatitis and psoriasis: a gene microarray
25. Malissen, B., Tamoutounour, S. & Henri, S. The origins and functions analysis. J. Allergy Clin. Immunol. 112, 1195–1202 (2003).
of dendritic cells and macrophages in the skin. Nat. Rev. Immunol. 55. Picardo, M. & Ottaviani, M. Skin microbiome and skin disease: the
14, 417–428 (2014). example of rosacea. J. Clin. Gastroenterol. 48(suppl. 1), S85–86 (2014).
26. Naik, S. et al. Compartmentalized control of skin immunity by 56. Hay, R.J. Malassezia, dandruff and seborrhoeic dermatitis. Br. J.
resident commensals. Science 337, 1115–1119 (2012). Dermatol. 165(suppl. 2), 2–8 (2011).
27. Scharschmidt, T.C. et al. A wave of regulatory T cells into neonatal 57. Verhulst, N.O. et al. Composition of human skin microbiota affects
skin mediates tolerance to commensal microbes. Immunity 43, attractiveness to malaria mosquitoes. PLoS One 6, e28991 (2011).
1011–1021 (2015). 58. Troccaz, M. et al. Mapping axillary microbiota responsible for body
28. Naik, S. et al. Commensal-dendritic-cell interaction specifies a unique odours using a culture-independent approach. Microbiome 3, 3 (2015).
protective skin immune signature. Nature 520, 104–108 (2015). 59. Egert, M. et al. rRNA-based profiling of bacteria in the axilla of healthy
29. Kong, H.H. Skin microbiome: genomics-based insights into the diversity males suggests right-left asymmetry in bacterial activity. FEMS
and role of skin microbes. Trends Mol. Med. 17, 320–328 (2011). Microbiol. Ecol. 77, 146–153 (2011).
30. Rosenthal, M., Goldberg, D., Aiello, A., Larson, E. & Foxman, B. Skin 60. Fujii, T., Shinozaki, J., Kajiura, T., Iwasaki, K. & Fudou, R. A newly
microbiota: Microbial community structure and its potential association discovered Anaerococcus strain responsible for axillary odor and a
with health and disease. Infect. Genet. Evol. 11, 839–848 (2011). new axillary odor inhibitor, pentagalloyl glucose. FEMS Microbiol. Ecol.
31. Probst, A.J. Auerbach, A.K. & Moissl-Eichinger, C. Archaea on human 89, 198–207 (2014).
skin. PLoS One 8, e65388 (2013). 61. Egert, M. et al. Identification of compounds inhibiting the C-S lyase
32. Nakatsuji, T. et al. The microbiome extends to subepidermal activity of a cell extract from a Staphylococcus sp. isolated from
compartments of normal skin. Nat. Commun. 4, 1431 (2013). human skin. Lett. Appl. Microbiol. 57, 534–539 (2013).
33. Gao, Z., Tseng, C., Pei, Z. & Blaser, M.J. Molecular analysis of human 62. Bawdon, D., Cox, D.S., Ashford, D., James, A.G. & Thomas, G.H.
forearm superficial skin bacterial biota. Proc. Natl. Acad. Sci. USA Identification of axillary Staphylococcus sp. involved in the production
104, 2927–2932 (2007). of the malodorous thioalcohol 3-methyl-3-sufanylhexan-1-ol. FEMS
34. Grice, E.A. et al. Topographical and temporal diversity of the human Microbiol. Lett. 362(16), fnv111 (2015).
skin microbiome. Science 324, 1190–1192 (2009). 63. Myles, I.A. et al. Transplantation of human skin microbiota in models of
35. Fierer, N., Hamady, M., Lauber, C.L. & Knight, R. The influence of atopic dermatitis. JCI Insight 1(10), e86955 (2016).
sex, handedness, and washing on the diversity of hand surface 64. Krutmann, J. Pre- and probiotics for human skin. J. Dermatol. Sci. 54,
bacteria. Proc. Natl. Acad. Sci. USA 105, 17994–17999 (2008). 1–5 (2009).
36. Findley, K. et al. Topographic diversity of fungal and bacterial 65. Zhang, M. et al. Oral antibiotic treatment induces skin microbiota dysbiosis
communities in human skin. Nature 498, 367–370 (2013). and influences wound healing. Microb. Ecol. 69, 415–421 (2015).
37. Egert, M. & Simmering, R. The microbiota of the human skin. Adv. 66. Scharschmidt, T.C. & Fischbach, M.A. What lives on our skin:
Exp. Med. Biol. 902, 61–81 (2016).
ecology, genomics and therapeutic opportunities of the skin
38. Human Microbiome Project Consortium. Structure, function and
microbiome. Drug Discov. Today Dis. Mech. 10, 3–4 (2013).
diversity of the healthy human microbiome. Nature 486, 207–214
67. Al-Ghazzewi, F.H. & Tester, R.F. Impact of prebiotics and probiotics
(2012).
on skin health. Benef. Microb. 5, 99–107 (2014).
39. McDonald, D. et al. Extreme Dysbiosis of the Microbiome in Critical
68. Grice, E.A. The skin microbiome: potential for novel diagnostic and
Illness (ed. Green Tringe, S.) mSphere. 1, e00199–16 (2016).
therapeutic approaches to cutaneous disease. Semin. Cutan. Med.
40. Christensen, G.J.M. & Bru €ggemann, H. Bacterial skin commensals and
Surg. 33, 98–103 (2014).
their role as host guardians. Benef. Microbes 5, 201–215 (2014).
69. Carolan, H., Watkins, S. & Bradshaw, D. The prebiotic concept—a
41. Iwase, T. et al. Staphylococcus epidermidis Esp inhibits
novel approach for skin health. Eur. Cosmet. 7/8, 22–27 (2008).
Staphylococcus aureus biofilm formation and nasal colonization.
70. Bockmu €hl, D.P. et al. Prebiotic cosmetics: an alternative to
Nature 465, 346–349 (2010).
antibacterial products. IFSCC Mag. 9, 197–200 (2006).
42. Zipperer, A. et al. Human commensals producing a novel antibiotic
71. Katsuyama, M. et al. A novel method to control the balance of skin
impair pathogen colonization. Nature 535, 511–516 (2016).
43. Belkaid, Y. & Segre, J.A. Dialogue between skin microbiota and microflora. Part 2. A study to assess the effect of a cream containing
immunity. Science 346, 954–959 (2014). farnesol and xylitol on atopic dry skin. J. Dermatol. Sci. 38, 207–213
44. Nakamizo, S. et al. Commensal bacteria and cutaneous immunity. (2005).
Semin. Immunopathol. 37, 73–80 (2015). 72. Donnarumma, G. et al. Malassezia furfur induces the expression of
45. Lai, Y. et al. Commensal bacteria regulate Toll-like receptor 3-dependent beta-defensin-2 in human keratinocytes in a protein kinase C-
inflammation after skin injury. Nat. Med. 15, 1377–1382 (2009). dependent manner. Arch. Dermatol. Res. 295, 474–481 (2004).
46. Lai, Y. et al. Activation of TLR2 by a small molecule produced by 73. Pernet, I. et al. An optimized method for intensive screening of
Staphylococcus epidermidis increases antimicrobial defense against molecules that stimulate beta-defensin 2 or 3 (hBD2 or hBD3)
bacterial skin infections. J. Invest. Dermatol. 130, 2211–2221 (2010). expression in cultured normal human keratinocytes. Int. J. Cosmet.
47. Belkaid, Y. & Naik, S. Compartmentalized and systemic control of Sci. 27, 161–170 (2005).
tissue immunity by commensals. Nat. Immunol. 14, 646–653 (2013). 74. Ouwehand, A.C., Båtsman, A. & Salminen, S. Probiotics for the skin:
48. Bru€ggemann, H. et al. The complete genome sequence of a new area of potential application? Lett. Appl. Microbiol. 36, 327–
Propionibacterium acnes, a commensal of human skin. Science 305, 331 (2003).
671–673 (2004). 75. Iovieno, A. et al. Preliminary evidence of the efficacy of probiotic eye-
49. Fitz-Gibbon, S. et al. Propionibacterium acnes strain populations in drop treatment in patients with vernal keratoconjunctivitis. Graefes
the human skin microbiome associated with acne. J. Invest. Arch. Clin. Exp. Ophthalmol. 246, 435–441 (2008).
Dermatol. 133, 2152–2160 (2013). 76. Mahe, Y.F. et al. Induction of the skin endogenous protective
50. Gao, Z., Tseng, C., Strober, B.E., Pei, Z. & Blaser, M.J. Substantial mitochondrial MnSOD by Vitreoscilla filiformis extract. Int. J. Cosmet.
alterations of the cutaneous bacterial biota in psoriatic lesions. PLoS Sci. 28, 277–287 (2006).
One 3, e2719 (2008). 77. Seite, S. et al. A new Vitreoscilla filiformis extract grown on spa water-
51. Alekseyenko, A.V. et al. Community differentiation of the cutaneous enriched medium activates endogenous cutaneous antioxidant and
microbiota in psoriasis. Microbiome 1, 31 (2013). antimicrobial defenses through a potential Toll-like receptor 2/
52. Kong, H.H. et al. Temporal shifts in the skin microbiome associated protein kinase C, zeta transduction pathway. Clin. Cosmet. Investig.
with disease flares and treatment in children with atopic dermatitis. Dermatol. 6, 191–196 (2013).
Genome Res. 22, 850–859 (2012). 78. Kimoto-Nira, H., Aoki, R., Sasaki, K., Suzuki, C. & Mizumachi, K. Oral
53. Kobayashi, T. et al. Dysbiosis and Staphylococcus aureus colonization intake of heat-killed cells of Lactococcus lactis strain H61 promotes
drives inflammation in atopic dermatitis. Immunity 42, 756–766 (2015). skin health in women. J. Nutr. Sci. 1, e18 (2012).

CLINICAL PHARMACOLOGY & THERAPEUTICS | VOLUME 102 NUMBER 1 | JULY 2017 69

You might also like