ArticlePDF AvailableLiterature Review

Abstract and Figures

Keloids constitute an abnormal fibroproliferative wound healing response in which raised scar tissue grows excessively and invasively beyond the original wound borders. This review provides a comprehensive overview of several important themes in keloid research: namely keloid histopathology, heterogeneity, pathogenesis, and model systems. Although keloidal collagen versus nodules and α-SMA-immunoreactivity have been considered pathognomonic for keloids versus hypertrophic scars, conflicting results have been reported which will be discussed together with other histopathological keloid characteristics. Importantly, histopathological keloid abnormalities are also present in the keloid epidermis. Heterogeneity between and within keloids exists which is often not considered when interpreting results and may explain discrepancies between studies. At least two distinct keloid phenotypes exist, the superficial-spreading/flat keloids and the bulging/raised keloids. Within keloids, the periphery is often seen as the actively growing margin compared to the more quiescent center, although the opposite has also been reported. Interestingly, the normal skin directly surrounding keloids also shows partial keloid characteristics. Keloids are most likely to occur after an inciting stimulus such as (minor and disproportionate) dermal injury or an inflammatory process (environmental factors) at a keloid-prone anatomical site (topological factors) in a genetically predisposed individual (patient-related factors). The specific cellular abnormalities these various patient, topological and environmental factors generate to ultimately result in keloid scar formation are discussed. Existing keloid models can largely be divided into in vivo and in vitro systems including a number of subdivisions: human/animal, explant/culture, homotypic/heterotypic culture, direct/indirect co-culture, and 3D/monolayer culture. As skin physiology, immunology and wound healing is markedly different in animals and since keloids are exclusive to humans, there is a need for relevant human in vitro models. Of these, the direct co-culture systems that generate full thickness keloid equivalents appear the most promising and will be key to further advance keloid research on its pathogenesis and thereby ultimately advance keloid treatment. Finally, the recent change in keloid nomenclature will be discussed, which has moved away from identifying keloids solely as abnormal scars with a purely cosmetic association toward understanding keloids for the fibroproliferative disorder that they are.
This content is subject to copyright.
fcell-08-00360 May 22, 2020 Time: 19:44 # 1
REVIEW
published: 26 May 2020
doi: 10.3389/fcell.2020.00360
Edited by:
Shiro Jimi,
Fukuoka University, Japan
Reviewed by:
Rajprasad Loganathan,
Johns Hopkins University,
United States
Lydia Masako Ferreira,
Federal University of São Paulo, Brazil
*Correspondence:
Susan Gibbs
s.gibbs@amsterdamumc.nl
Specialty section:
This article was submitted to
Cell Growth and Division,
a section of the journal
Frontiers in Cell and Developmental
Biology
Received: 02 December 2019
Accepted: 22 April 2020
Published: 26 May 2020
Citation:
Limandjaja GC, Niessen FB,
Scheper RJ and Gibbs S (2020) The
Keloid Disorder: Heterogeneity,
Histopathology, Mechanisms
and Models.
Front. Cell Dev. Biol. 8:360.
doi: 10.3389/fcell.2020.00360
The Keloid Disorder: Heterogeneity,
Histopathology, Mechanisms and
Models
Grace C. Limandjaja1, Frank B. Niessen2, Rik J. Scheper3and Susan Gibbs1,4*
1Department of Molecular Cell Biology and Immunology, Amsterdam University Medical Center (location VUmc), Vrije
Universiteit Amsterdam, Amsterdam, Netherlands, 2Department of Plastic Surgery, Amsterdam University Medical Center
(location VUmc), Vrije Universiteit Amsterdam, Amsterdam, Netherlands, 3Department of Pathology, Amsterdam University
Medical Center (location VUmc), Vrije Universiteit Amsterdam, Amsterdam, Netherlands, 4Department of Oral Cell Biology,
Academic Centre for Dentistry (ACTA), University of Amsterdam and Vrije Universiteit Amsterdam, Amsterdam, Netherlands
Keloids constitute an abnormal fibroproliferative wound healing response in which
raised scar tissue grows excessively and invasively beyond the original wound borders.
This review provides a comprehensive overview of several important themes in
keloid research: namely keloid histopathology, heterogeneity, pathogenesis, and model
systems. Although keloidal collagen versus nodules and α-SMA-immunoreactivity have
been considered pathognomonic for keloids versus hypertrophic scars, conflicting
results have been reported which will be discussed together with other histopathological
keloid characteristics. Importantly, histopathological keloid abnormalities are also
present in the keloid epidermis. Heterogeneity between and within keloids exists which is
often not considered when interpreting results and may explain discrepancies between
studies. At least two distinct keloid phenotypes exist, the superficial-spreading/flat
keloids and the bulging/raised keloids. Within keloids, the periphery is often seen as
the actively growing margin compared to the more quiescent center, although the
opposite has also been reported. Interestingly, the normal skin directly surrounding
keloids also shows partial keloid characteristics. Keloids are most likely to occur after an
inciting stimulus such as (minor and disproportionate) dermal injury or an inflammatory
process (environmental factors) at a keloid-prone anatomical site (topological factors)
in a genetically predisposed individual (patient-related factors). The specific cellular
abnormalities these various patient, topological and environmental factors generate
to ultimately result in keloid scar formation are discussed. Existing keloid models can
largely be divided into in vivo and in vitro systems including a number of subdivisions:
human/animal, explant/culture, homotypic/heterotypic culture, direct/indirect co-culture,
and 3D/monolayer culture. As skin physiology, immunology and wound healing is
markedly different in animals and since keloids are exclusive to humans, there is a
need for relevant human in vitro models. Of these, the direct co-culture systems that
generate full thickness keloid equivalents appear the most promising and will be key
to further advance keloid research on its pathogenesis and thereby ultimately advance
Frontiers in Cell and Developmental Biology | www.frontiersin.org 1May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 2
Limandjaja et al. The Keloid Disorder
keloid treatment. Finally, the recent change in keloid nomenclature will be discussed,
which has moved away from identifying keloids solely as abnormal scars with a purely
cosmetic association toward understanding keloids for the fibroproliferative disorder that
they are.
Keywords: keloid, cicatrix, hypertrophic, keloid anatomy and histology, keloid etiology, keloid pathology, keloid
heterogeneity, keloid model, keloid phenotype
INTRODUCTION
As early as approximately 3000 B.C., the existence of keloid scars
has been acknowledged in the description of a swelling on his
breast, large, spreading, and hard, which felt like touching a
ball of wrappings in the Edwin Smith Papyrus, the first known
surgical treatise describing ancient Egyptian medical practice
(Breasted, 1930;Berstein and Roenigk, 1996). Keloids are not
mentioned in modern day literature until the early 19th century
when Jean Louis Alibert, the father of French Dermatology,
first described tumor-like scars which he initially referred to
as les cancroïdes de la peau. When it became clear these
cicatricial tumors were in fact non-cancerous, Alibert changed
the name to cheloïde or keloïde in reference to the Greek word
χηλi (khçlçé) for crab’s claw and the suffix -oid meaning ‘like.’
Together this is meant to reflect not only the claw-like extension
of the keloids but also refers to their horizontal invasive growth
beyond the initial wound margins into the surrounding skin
(Alibert, 1825;Delpech, 1881;Berstein and Roenigk, 1996).
The keloid incidence rate varies greatly and is known to be
influenced by racial ethnicity. The risk of keloid development
significantly increases with increasing pigmentation (Burd and
Huang, 2005;Wolfram et al., 2009). In the Black and the Hispanic
general population, the incidence varies from 4.5–6.2 to 16%
(Cosman et al., 1961;Oluwasanmi, 1974;Rockwell et al., 1998);
while the incidence in the Taiwanese Chinese and Caucasians is
reported to be as low as <1% (Bloom, 1956;Seifert and Mrowietz,
2009;Sun et al., 2014). However, these numbers are largely based
on studies from several decades ago with the oldest dating back
to 1931. To our knowledge there are no new incidence numbers
of keloid scarring in the general population. More recent data
is available for very specific subpopulations: in head and neck
surgical patients (Young et al., 2014) as well as women after
caesarian section (Tulandi et al., 2011), the incidence of keloid
scar formation was significantly increased in African Americans
(0.8 and 7.1%, respectively) compared with the Caucasian (0.1
and 0.5%, respectively) and Asian or other (0.2 and 5.2%,
respectively) population. Interestingly, in Africans with albinism
the prevalence rate of 7.5% was not statistically different from
the overall prevalence rate of 8.3% in the general population or
the 8.5% observed in the normally pigmented African population
(Kiprono et al., 2015). It would therefore seem that increased
pigmentation in and of itself cannot solely explain the reported
ethnic differences in incidence rates (Bran et al., 2009).
In addition to the obvious cosmetic disfigurement, keloids
can also produce symptoms of itching and pain (Lee S. S. et al.,
2004;Bijlard et al., 2017). A study comparing the quality of life
in patients with keloids to that of psoriasis patients found that
patients with abnormal scars demonstrated the same reduced
quality of life levels as psoriasis patients when compared with
healthy controls (Balci et al., 2009). Similarly, a cross-sectional
survey on the burden of keloid disease (Bijlard et al., 2017)
showed that having keloids was associated with considerable
impairment of emotional wellbeing. In summary, keloids may
affect a very specific demographic for reasons we do not yet know,
but for those affected, these abnormal scars can have significant
consequences beyond cosmetics.
The mechanisms behind keloid scarring in particular are still
poorly understood (Slemp and Kirschner, 2006;Robles and Berg,
2007;Seifert and Mrowietz, 2009), and this is reflected in our
inability to satisfactorily manage this abnormal scar (Niessen
et al., 1999;Butler et al., 2008b). Known for its therapy-resistant
nature, excision alone has recurrence rates of 55–100% and can
even result in the development of a worse scar than before (Robles
and Berg, 2007;Butler et al., 2008b;Balci et al., 2009;Shih
et al., 2010). To further advance research on the pathogenesis
underlying keloid scar formation, there is an urgent need for
relevant, true-to-life keloid scar models that resemble the in vivo
keloid phenotype accurately. This review on keloid scars will
discuss histopathological characteristics, inter- and intralesional
heterogeneity, the pathogenetic mechanisms, as well as existing
scar model systems of keloids.
KELOID HISTOPATHOLOGY
Keloids are primarily a clinical diagnosis (Gulamhuseinwala
et al., 2008), and as such are not usually sent in for further
analysis by the pathologist. Although the histopathological
definition of a keloid scar was not further detailed in the
article, Gulamhuseinwala et al. (2008) found that retrospective
analysis of H&E stainings of 568 clinically diagnosed keloids
only proved accurate in 81% of the cases. Experienced plastic
surgeons diagnosed keloids based on the following clinical
criteria: the presence of a scar with a history of antecedent local
trauma and growth extending beyond its boundary. The non-
keloid diagnoses included acne keloidalis (11%), hypertrophic
(6%), and even normotrophic (2%) scars and a single pilonidal
abscess. Importantly though, no malignancies or dysplasias were
reported. Based on these findings, the authors suggested that
sending excised keloid tissue for histopathological examination
is not necessary if the clinician is an expert and there is a
strong clinical suspicion (Gulamhuseinwala et al., 2008). In
response to this study, however, Wong and Ogawa pointed
out that many clinicians would not be comfortable with the
incorrect diagnosis rate of 19% and therefore advocate for
Frontiers in Cell and Developmental Biology | www.frontiersin.org 2May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 3
Limandjaja et al. The Keloid Disorder
post-surgical histopathological confirmation (Wong and Lee,
2008;Ogawa et al., 2009).
The histopathological abnormalities of the full scarring
spectrum and normal skin have been summarized in
Supplementary Table S1, specific cellular abnormalities in
keloid scars are summarized in Supplementary Table S3 and will
be elaborated upon in the section ‘Keloid cellular abnormalities’.
The histopathological findings on keloid scars will be briefly
summarized in this section. The epidermal thickness in keloid
scars has been described as anything from atrophic (Koonin,
1964;Bakry et al., 2014) and normal (Moshref and Mufti, 2009;
Huang et al., 2014), to sometimes (Ehrlich et al., 1994;Materazzi
et al., 2007) or always increased (Bertheim and Hellström, 1994;
Chua et al., 2011;Syed et al., 2011;Sidgwick et al., 2013;Jumper
et al., 2015;Suttho et al., 2017;Shang et al., 2018). However, the
overwhelming majority supports the observation of increased
epidermal thickness in keloid scars, and what is more, this was
confirmed when thickness was measured in µm (Hellström et al.,
2014) as well as number of viable cell layers (Limandjaja et al.,
2017, 2019). Similarly, conflicting findings have been reported
with regards to rete ridge formation. Reports range from normal
rete ridge formation (Lee J. Y. Y.et al., 2004;Moshref and Mufti,
2009) to reduced (Koonin, 1964;Chong et al., 2015;Jumper
et al., 2015;Suttho et al., 2017;Shang et al., 2018) or complete
absence thereof (Ehrlich et al., 1994;Meenakshi et al., 2005;
Huang et al., 2014), although none have attempted to objectively
measure the extent of rete ridge formation. Overall, most
studies, including our own histopathological studies (Limandjaja
et al., 2017, 2019), appear to support the findings of a flattened
epidermis with increased thickness. Epidermal differentiation
appears mostly unaffected (Bloor et al., 2003;Ong et al., 2010;
Limandjaja et al., 2017, 2019). Although increased epidermal
activation/proliferation has been observed (Prathiba et al., 2001;
Bloor et al., 2003;Ong et al., 2010), this is in contrast with the
findings from our extensive immunohistochemical analysis
of the keloid epidermis (Limandjaja et al., 2017). We showed
normal levels of epidermal proliferation and differentiation in the
thickened keloid epidermis, save for the precocious expression
of terminal differentiation marker involucrin. We therefore
proposed that increased epidermal thickness was not the result of
epidermal hyperproliferation, but was associated with abnormal
differentiation instead. Interestingly, keloid keratinocytes
also showed increased expression of epithelial-mesenchymal
transition (EMT) markers (Chua et al., 2011;Ma et al., 2015;Yan
et al., 2015;Hahn et al., 2016;Kuwahara et al., 2016).
Histopathological studies of the keloid dermis showed that
fibroblasts were present in higher numbers (Ueda et al., 1999;
Tanaka et al., 2004;Meenakshi et al., 2009;Jiao et al., 2017).
Other dermal cell types residing in keloids include myofibroblasts
(Santucci et al., 2001;Kamath et al., 2002;Lee J. Y. Y.et al.,
2004; Lee et al., 2012;Moshref and Mufti, 2009;Shin et al., 2016)
and fibrocytes (Iqbal et al., 2012;Shin et al., 2016), both were
present in increased numbers. In our whole biopsy image analysis
of keloid tissue, CD34 expression was found to be absent from
the keloid dermis, but constitutively and abundantly present in
normal skin and normotrophic scars (Limandjaja et al., 2019).
Interestingly, within these CD34dermal regions, we found
senescent (p16+) mesenchymal cells (vimentin+) as well as
myofibroblasts (α-SMA+).
Overall trends in the dermal ECM composition include
increased levels of collagen I and III (Naitoh et al., 2001;Syed
et al., 2011) with increased collagen bundle thickness (Verhaegen
et al., 2009); increased fibronectin (Kischer and Hendrix, 1983),
glycosaminoglycans (Carrino et al., 2012), chondroitin sulfate
(Ikeda et al., 2009), biglycan (Hunzelmann et al., 1996), versican
(Carrino et al., 2012), tenascin (Dalkowski et al., 1999), and
periostin (Zhou et al., 2010;Maeda et al., 2019); while levels
of elastin (Szulgit et al., 2002;Ikeda et al., 2009;Theoret et al.,
2013), and decorin (Carrino et al., 2012) were reduced. Dermal
hyaluronic acid expression showed variable results (Alaish et al.,
1995;Meyer et al., 2000;Ikeda et al., 2009;Yagi et al., 2013), but
unlike normal skin, its expression was equal in both the keloid
epidermis and dermis (Tan et al., 2011). Reports on vascularity
are also highly variable, both increased (Ehrlich et al., 1994;
Amadeu et al., 2003;Tanaka et al., 2004;Materazzi et al., 2007;
Ong et al., 2007b;Syed and Bayat, 2012;Bakry et al., 2014) and
decreased vascular density (Beer et al., 1998;Ueda et al., 2004;
Kurokawa et al., 2010;Theoret et al., 2013) has been observed
in keloids. Nerve fiber density appears to be increased in keloids
compared with normal skin (Hochman et al., 2008;Drummond
et al., 2017). Lastly, keloids also show increased immune cell
infiltration (Amadeu et al., 2003;Sharquie and Al-Dhalimi, 2003;
Tanaka et al., 2004;Shaker et al., 2011;Jiao et al., 2015;Luo et al.,
2017), with higher quantities of macrophages (Boyce et al., 2001;
Shaker et al., 2011;Jiao et al., 2015) and T-lymphocytes (Shaker
et al., 2011;Murao et al., 2014;Jiao et al., 2015) in particular. An
extensive review by Jumper et al. (2015) on the histopathology
of keloid scars has reiterated most of the aforementioned
findings, and further emphasizes the following as most frequently
occurring and therefore most discerning features of keloid scars:
a thickened, flattened epidermis; a tongue-like advancing edge
in the dermis; haphazard, thick, hyalinized collagen bundles as
the predominant dermal feature, with subsequent loss of the
papillary-reticular boundary; increased dermal cellularity; signs
of inflammation; and variable α-SMA expression.
Unlike their keloid counterparts, hypertrophic scars are
raised scars whose growth remains within the borders of the
original wound (Burd and Huang, 2005) and they can be
difficult to distinguish histopathologically. Keloidal collagen
(Cosman et al., 1961;Santucci et al., 2001;Ogawa et al.,
2009) vs. α-SMA and dermal nodules (Ehrlich et al., 1994;
Huang et al., 2014) have often been cited as pathognomonic
features for keloids or hypertrophic scars, respectively, but
conflicting reports abound (Muir, 1990;Ehrlich et al., 1994;
Santucci et al., 2001;Lee J. Y. Y.et al., 2004;Ogawa et al.,
2009;Ali et al., 2010;Bux and Madaree, 2010;Huang et al.,
2014). In a histopathological study (Limandjaja et al., 2019)
comparing both scar types, α-SMA and dermal nodules were
present in both scars and while keloidal collagen remained
a strong keloid marker, it was also observed in one of the
hypertrophic scars. Additionally, a thickened epidermis with
involucrin overexpression and a CD34/α-SMA+/p16+dermal
cell population could be found in both scar types, although
α-SMA and p16 immunoreactivity were present in higher degrees
Frontiers in Cell and Developmental Biology | www.frontiersin.org 3May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 4
Limandjaja et al. The Keloid Disorder
in hypertrophic scars vs. keloids, respectively. In short, despite
the clearly defined clinical distinction between the two abnormal
scar types, the histological distinction between hypertrophic and
keloid scars remains a source of contention, especially in the early
stages (Burd and Huang, 2005).
INTER- AND INTRALESIONAL KELOID
HETEROGENEITY
Interlesional Heterogeneity
Several reports suggest that distinct keloid phenotypes may exist.
As early as 1960, Conway et al. (1960) discerned between nodular
raised keloids and flat keloids often observed on the sternum.
More recently, Bella et al. (2011) studied large multigenerational
pedigrees of patients with familial keloids in three rural African
tribes. The superficial spreading phenotype predominated in two
of the tribes, while the raised vertical phenotype predominated
in the remaining tribe (Figure 1). Superficial spreading keloids
show irregular subepidermal spread with irregular areas of hyper-
and hypopigmentation, they are mostly raised at the edges and
are characterized by a central flattened and quiescent area. This
central area is often regularly or hypopigmented, while the
margins show hyperpigmentation. In contrast, raised keloid scars
are prominently bulbous in shape with distinct borders and
may have limited areas of central quiescence. Another form of
interlesional keloid heterogeneity was proposed by Akaishi et al.
(2010), who distinguished between regular keloids with a round
shape and clear curving lines, and irregular keloids with irregular
shapes and lines. The authors found that the irregularly shaped
keloids showed a significant increase in infection and previous
surgery rates, and proposed that in contrast, the shape of regular
keloids was determined by skin tension alone. In predilection
body sites exposed to constant stretching (e.g., scapula, chest, and
shoulder), the butterfly, the crab’s claw or the dumbbell have also
been described as typical keloid shapes which are predominantly
determined by local mechanical factors (Huang et al., 2017). In
short, several distinct keloid phenotypes have been described,
the division of which appears to be based predominantly on the
keloid’s growth pattern and/or the resulting shape.
Intralesional Heterogeneity
It is also likely that heterogeneity exists within keloid scars
(see Supplementary Table S1). Based on clinical observations,
the most often described distinction is that of a red, raised
peripheral margin which actively invades the surrounding skin
and more depressed, lighter colored center showing clinical
regression (Louw et al., 1997;Lu et al., 2007a;Seifert et al., 2008).
This peripheral-central distinction matches the description of
the superficial spreading keloid phenotype (Bella et al., 2011).
Symptoms of strong itching (Lee S. S. et al., 2004) predominate
in the more pigmented keloid margin (Louw et al., 1997;
Le et al., 2004;Lu et al., 2007a;Bella et al., 2011), together
with hypercellularity (Appleton et al., 1996;Ladin et al., 1998;
Akasaka et al., 2001;Varmeh et al., 2011;Huang et al., 2014),
increased vascularity (Appleton et al., 1996;Le et al., 2004;Touchi
et al., 2016) and immune cell infiltration (Appleton et al., 1996;
FIGURE 1 | Spreading vs. bulging keloid phenotypes. Watercolor illustration.
Left figure shows a typical keloid of the ‘spreading’ phenotype located on the
anterior chest, with quiescent center and an actively growing peripheral
margin. Right figure shows a keloid of the ‘bulging’ phenotype, which are
bulbous in shape and can often be observed on the earlobe. Figure first
published in Limandjaja (2019), used with permission.
Le et al., 2004). Reduced apoptosis (Lu et al., 2007a;Seifert et al.,
2008), increased proliferation (Varmeh et al., 2011;Suttho et al.,
2017) and increased cellular activity (Louw et al., 1997), all
contribute to enlarging the pool of ECM-producing fibroblasts in
this region and support the hypothesized increased keloid activity
(see Supplementary Table S2A). In contrast, the central keloid
region shows either hypopigmentation or regular pigmentation
(Louw et al., 1997;Le et al., 2004;Lu et al., 2007a;Bella et al.,
2011) with pain as the main symptom (Lee S. S. et al., 2004), as
well as hypocellularity (Appleton et al., 1996;Ladin et al., 1998;
Akasaka et al., 2001;Varmeh et al., 2011;Huang et al., 2014)
and reduced vascularity (Appleton et al., 1996;Le et al., 2004;
Touchi et al., 2016) (see Supplementary Table S2A). Fibroblasts
derived from this central region generally show signs of inactivity
(Louw et al., 1997), as well as reduced proliferation (Varmeh et al.,
2011;Suttho et al., 2017), increased apoptosis (Lu et al., 2007a;
Seifert et al., 2008) and senescence (Varmeh et al., 2011). Taken
together with the increased expression of ECM-degrading genes
(Seifert et al., 2008), the central region appears to be the area of
relative quiescence.
Overall, the majority of studies support the concept of an
actively developing periphery and a more quiescent central
region, but the reverse has also been postulated with an active
role for the central region rather than the periphery. The
keloid center has been reported to show increased proliferation
(Giugliano et al., 2003;Tsujita-Kyutoku et al., 2005), the absence
of apoptosis (Appleton et al., 1996;Sayah et al., 1999;Akasaka
et al., 2001), increased expression of fibrosis-associated genes
(e.g., TGFβRI, SMAD 2, and SMAD3) (Tsujita-Kyutoku et al.,
2005) and certain wound healing mediators (IL-6 and VEGF)
(Giugliano et al., 2003), all of which support a more pro-active
role for this region. In line with these findings, our in vitro
reconstructed (Limandjaja et al., 2018b) different keloid regions
showed that differences existed between the regions in terms
of scar parameter expression. The central deep keloid region
Frontiers in Cell and Developmental Biology | www.frontiersin.org 4May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 5
Limandjaja et al. The Keloid Disorder
showed the more exaggerated keloid phenotype with respect to
increased contraction, increased epidermal thickness, reduced
HGF secretion and reduced collagen type IV α2 chain dermal
gene expression.
Comparison of the keloid heterogeneity findings remains
difficult due to the varying definitions of what constitutes the
periphery and the center within a keloid, as this may differ
significantly between studies. Additionally, we suspected that the
apparent dichotomy between results supporting an active keloid
periphery and those whose findings ascribe this active role to
the keloid center, may be explained by the existence of different
keloid phenotypes (Limandjaja et al., 2018b). Various keloid
phenotypes have been described in the previous paragraph, but
we would like to propose an even simpler classification based on
the work of Bella et al. (2011) and Supp et al. (2012b): namely that
of a more concave ‘superficial spreading and raised or ‘bulging’
keloids (see Figure 1) (Bella et al., 2011). Depending on the
phenotype, the actively expanding region could be located in
the periphery or the deeper central region (Supp et al., 2012b).
As all the keloids included in the study were of the ‘bulging’
phenotype, it follows that the central deep region would show the
most aggressive keloid phenotype.
Furthermore, heterogeneity has also been observed in regions
other than the periphery and the center, these are summarized
in Supplementary Table S2B. An often-used division is that
between different dermal layers (Russell et al., 1989, 1995;Luo
et al., 2001;Supp et al., 2012b;Chong et al., 2015, 2018;
Jiao et al., 2017), in which case the middle or deepest dermal
layers were usually found to act more aggressively compared
with the more superficial layers. The conflicting reports on the
peripheral and central keloid regions notwithstanding, it is clear
that differences exist between different lesional sites within a
keloid and these abnormal scars should therefore not simply
be considered or treated as a homogenous growth. Because of
this, we strongly advocate for the inclusion of a description of
the keloid phenotype/shape and the use of schematic drawings
to indicate from where within a keloid samples were taken
for experimentation.
Normal Skin Surrounding Keloids
The normal skin directly adjacent to and surrounding keloid scars
is also rarely included in keloid research. Although similarity to
normal skin has been reported (Jumper et al., 2017), the majority
of reported results suggest that the surrounding normal skin
behaves more like keloid tissue than normal skin. For example,
the surrounding normal skin often itches like the keloid periphery
(Lee S. S. et al., 2004) and shows increased blood flow compared
with unaffected normal skin (Liu et al., 2016). On a cellular level,
increased staining of the hematopoietic stem cell marker c-KIT
(Bakry et al., 2014) and heat shock protein 70 (Lee et al., 2013) has
been observed in both keloids and their surrounding normal skin,
while keratinocytes and fibroblasts from the surrounding normal
skin shared the abnormal expression of many of the same genes in
keloid-derived keratinocytes and fibroblasts (Hahn et al., 2013).
In the dermal compartment, the epidermal appendages lost from
keloid tissue reappear in the surrounding normal skin in reduced
capacity; and the dense and excessive collagen deposition of the
keloid can extend into the ECM of the surrounding normal skin,
which is otherwise loosely organized, with thin, wavy or even
fragmented collagen fibers (Lee et al., 2013; Lee W. J. et al., 2015;
Bakry et al., 2014;Jiao et al., 2017). Similarly, portions of nodules
from the adjacent keloid have also been found to extend into the
surrounding normal skin (Kischer and Pindur, 1990). The dermis
adjacent to keloids is more cellular but less crowded compared
with healthy skin, and shows significant lymphocyte infiltration
(Bakry et al., 2014;Lee W. J. et al., 2015;Jiao et al., 2017). The
skin surrounding keloids also differed from unaffected healthy
skin with respect to proliferation and apoptosis. Increased dermal
proliferation and increased numbers of apoptotic keratinocytes
(Appleton et al., 1996) were only observed in the surrounding
normal skin, and absent from healthy skin. Lastly, Dohi et al.
(2019) proposed an even more prominent role for surrounding
normal skin as the driving force for keloid progression into
the normal skin, via local preferential increased mechanical
strain. Conversely, the surrounding normal skin has also been
reported to differ from keloids. Compared with keloid scars, the
surrounding normal skin has more blood vessels (Beer et al.,
1998) and showed increased expression of the proliferative PCNA
marker in the dermis, which was normally absent in both normal
skin and keloids (Appleton et al., 1996). The surrounding skin
also showed strong, increased levels of CD34 staining in contrast
with the CD34 absence in abnormal scars (Erdag et al., 2008).
It also lacks the abnormal thermosensory thresholds to warmth,
cold, heat and pain sensations reported for keloids (Lee S. S.
et al., 2004). All things considered, current literature suggests
that the surrounding normal skin shares several features with
the adjacent keloid and is therefore a relevant area to include for
further investigation.
In our histopathological analysis of keloids, we found that the
normal skin directly adjacent to the keloids mostly resembled
normal skin or mature normotrophic scar (Limandjaja et al.,
2019). An epidermis of normal thickness and rete ridge formation
with normal differentiation and proliferation could be seen,
together with a CD34+/α-SMA/p16phenotype instead of the
CD34/α-SMA+/p16+phenotype associated with keloid scars.
However, the most interesting findings emerged from our in vitro
work (Limandjaja et al., 2018b), where changes were observed
which in and of itself were not statistically significant but overall
formed a clear pattern of intermediate abnormal expression.
Across all the abnormal scar parameters present in the in vitro
keloid scar model, the in vitro reconstructed surrounding normal
skin showed a phenotype more extreme than true normal skin but
less aggressive than the peripheral keloid models. With respect
to contraction, α-SMA immunoreactivity, HGF secretion and
collagen type IV α2 gene expression, the surrounding normal
skin showed intermediate between truly unaffected normal skin
and keloid scar.
Taken together, we and others have shown that heterogeneity
exists within keloid scars. For future studies it would therefore
be imperative to mention the shape and growth pattern of the
keloid (superficial spreading or bulging) and additionally, it
should always be mentioned where in the keloid any tissue for
experimentation was obtained from, preferably with a schematic
overview for unambiguous clarification. Additionally, we would
Frontiers in Cell and Developmental Biology | www.frontiersin.org 5May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 6
Limandjaja et al. The Keloid Disorder
argue that it is worth including the surrounding normal skin in
any keloid research study when possible.
KELOID SCAR PATHOGENESIS: WHERE
DO WE STAND?
The development of a human in vitro keloid scar model
resembling in vivo keloid tissue would not only benefit drug
development and testing, it would also greatly aid research into
the underlying mechanisms leading to keloid scarring. The fact
that keloid pathogenesis remains so poorly understood (Slemp
and Kirschner, 2006;Wolfram et al., 2009), has been the bane of
affected patients and clinicians alike. Despite the many theories
proposed by experts, no single unifying hypothesis has been
put forward (Seifert and Mrowietz, 2009). In Figure 2, adapted
from Wolfram et al. (2009), the various abnormalities reported
in keloid tissue and keloid-derived cells are organized in several
tiers: patient, topography or special skin sites, and environmental
factors all contribute to the expression of abnormal cellular
responses, which eventually lead to keloid scar formation. In the
following sections, recent and relevant findings for each of these
factors will be discussed.
Patient-Related Factors Influencing
Keloid Scarring
Ethnicity, genetic predisposition, gender and age are patient
characteristics which may influence keloid predilection (Wolfram
et al., 2009). There is no conclusive evidence in favor of
differences in occurrence based on gender. Some have reported
that keloids are more likely to occur in women than men (Bayat
et al., 2005;Burd and Huang, 2005;Seifert and Mrowietz, 2009;
Middelkoop et al., 2011;Young et al., 2014), but this may also
reflect at least in part, the overall greater awareness of unaesthetic
scarring in women and a consequent higher tendency to seek
medical assistance (Marneros et al., 2001;Burd, 2006). Young
age is also associated with increased risk of abnormal scarring
(Middelkoop et al., 2011). Keloids can develop at any age, but
incidence is highest between the ages of 10–30 years (Bayat
et al., 2005;Seifert and Mrowietz, 2009;Young et al., 2014;Lu
et al., 2015). Because of the peak in incidence immediately post-
puberty, exacerbations during pregnancy and resolution after
menopause, a potential role for endocrinological hyperactivity
in keloid pathogenesis has also been proposed (Rockwell et al.,
1998;Seifert and Mrowietz, 2009;Huang and Ogawa, 2013a;
Glass, 2017).
Perhaps the most relevant patient-related factor is the genetic
predisposition for keloid formation in certain individuals.
Ethnic differences in prevalence showed that darker pigmented
individuals are at higher risk of developing keloid scars
(Marneros et al., 2001;Al-Attar et al., 2006). Additionally, having
a family member with keloids is associated with increased keloid
prevalence (Bayat et al., 2005;Kiprono et al., 2015;Lu et al.,
2015). This is predominantly the case for first degree relatives,
as demonstrated by a heritability of 72, 41, and 17% for first,
second, and third degree relatives in the Chinese population
(Lu et al., 2015). Individuals with a family history of keloids
are also at higher risk of developing multiple keloids and
developing keloids of greater severity (Bayat et al., 2005;Lu
et al., 2015). The familial heritability, the increased prevalence
in certain ethnicities and common occurrence in twins, all
strongly support the concept of genetic susceptibility in patients
with keloid scars (Marneros et al., 2001;Shih and Bayat, 2010;
Glass, 2017). Other lines of evidence pointing to a genetic
influence on keloid predisposition include familial inheritance
patterns, linkage studies, case-control association studies, and
gene expression studies (Shih and Bayat, 2010;Glass, 2017).
Different modes of inheritance have been reported, varying
from autosomal recessive and X-linked to autosomal dominant
(Shih and Bayat, 2010;Glass, 2017). Shih and Bayat (2010)
have previously reviewed the available evidence and suggest that
most evidence points to an autosomal dominant inheritance
pattern with incomplete penetration and variable expression,
this then also explains why carriers do not always express
the keloid phenotype and why keloid patients do not always
respond to trauma with keloid scarring. Despite their valuable
contribution to our understanding of genetic predilection for
keloids, familial inheritance studies have not led to the discovery
of any particular predisposing genes (Glass, 2017). Multiple gene
mapping methods as well as targeted gene pathway investigations
have identified several gene polymorphisms (NEDD4, FOXL2,
MYO1E, and MYO7A also HLA) associated with keloids
(Nakashima et al., 2010;Zhu et al., 2013;Velez Edwards
et al., 2014;Glass, 2017), but the underlying mechanism is
still unclear. Similarly, various abnormalities in gene expression
have shown highly variable results between studies (Shih and
Bayat, 2010), but affected genes are known to be involved in
the ECM, inflammation and apoptosis (Shih and Bayat, 2010).
Aside from these inherited gene mutations, acquired altered
gene expression in the form of epigenetic modification may
also play a role in keloid pathogenesis and further complicates
matters (Glass, 2017;He et al., 2017). Ultimately, however, the
specific genetic variation responsible for keloid scarring has yet
to be identified, but likely involves more than a single gene.
Additionally, different keloid patients probably carry different
gene polymorphisms which can all lead to keloid scar formation,
this would explain the variations in keloid phenotype observed in
different people (Shih and Bayat, 2010;Velez Edwards et al., 2014;
Glass, 2017).
Topography-Related Factors Influencing
Keloid Scarring
It is important to note that patients with a history of keloid
scarring do not necessarily form keloids after every injury (Slemp
and Kirschner, 2006), two identical incisions can generate one
normal scar and one keloid in the same individual (Fong et al.,
1999;Fong and Bay, 2002;Al-Attar et al., 2006). Certain body
sites are more prone to keloid scarring, thus the location of
the wound influences risk of keloid scar formation (Wolfram
et al., 2009;Middelkoop et al., 2011). The earlobe, neck, sternum,
upper back, shoulders and upper limbs all constitute keloid-
prone anatomical sites (Murray et al., 1981;Bayat et al., 2005;
Burd and Huang, 2005;Bella et al., 2011;Middelkoop et al., 2011).
Frontiers in Cell and Developmental Biology | www.frontiersin.org 6May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 7
Limandjaja et al. The Keloid Disorder
FIGURE 2 | Pathogenesis of keloid scarring. Overview of the various factors involved in keloid pathogenesis, adapted from Wolfram et al. (2009). This figure is meant
to provide a provisionary framework to help organize the multitude of pathogenetic findings in a logical and systematic way. Abbreviations; ECM, extracellular matrix;
EMT, epithelial-mesenchymal transition; EndoMT, endothelial-mesenchymal transition; mϕ, macrophage; M2, alternatively activated, pro-fibrotic macrophage
subtype; PBMCs, peripheral blood mononuclear cells. Figure first published in Limandjaja (2019), used with permission.
Although the keloid-prone earlobe and anterior chest have
been described as tension-free areas (Brody, 1990;Ogawa et al.,
2003;Al-Attar et al., 2006), the most popular explanation for why
keloids occur more frequently at certain body sites remains the
theory that these are regions of increased skin tension that are
subject to constant stretching during normal movement (Peacock
et al., 1970;Rockwell et al., 1998;Butler et al., 2008a;Bux and
Madaree, 2012;Ogawa et al., 2012). There is no consensus on
whether elasticity may explain the differences between keloid-
prone and keloid-protected sites. Bux and Madaree (2012)
reported that keloid-prone sites are characterized by high tension
with low stretch and low elastic modulus. In contrast, Sano
et al. (2018) observed that with exception of the earlobes, sites
less prone to pathologic scarring (e.g., palmoplantar regions)
were comparatively hard, characterized by low distensibility
and reduced elasticity. In contrast, keloid susceptible sites showed
high distensibility and increased elasticity. Additionally, high
sebaceous gland density (Yagi et al., 1979;Fong et al., 1999;Fong
and Bay, 2002;Al-Attar et al., 2006), increased collagen and
decreased M1 macrophage numbers (Butzelaar et al., 2017), are
all characteristics of keloid-prone skin which may promote keloid
scar formation in genetically predisposed individuals.
Environmental Factors Influencing
Keloid Scarring
Although spontaneous keloid scar formation has been reported
(Jfri et al., 2015), it is a rare occurrence that has been reported
in association with certain syndromes such as Rubenstein-Taybi
and Goeminne syndrome (Jfri et al., 2015), or may simply
reflect a laps in memory (Robles and Berg, 2007;Jfri et al.,
2015). Environmental factors are therefore generally an essential
prerequisite to keloid scar formation as some form of assault to
the skin has to take place to incite keloidogenesis (Bran et al.,
2009;Wolfram et al., 2009;Shih and Bayat, 2010). These keloid-
inducing events vary from minor to major antecedent trauma,
as well as any process resulting in skin inflammation. Insect
bites or vaccinations are examples of minor insults to the skin
which may be so minor as to not be remembered by the patient
at all, while major trauma is usually observed in the setting
Frontiers in Cell and Developmental Biology | www.frontiersin.org 7May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 8
Limandjaja et al. The Keloid Disorder
of surgical and non-surgical wound healing. Examples of the
latter group include lacerations, abrasions, piercings, tattooing
or blunt trauma. Additionally, inflammatory skin conditions
such as acne, (peri)folliculitis, chicken pox, herpes zoster and
hidradenitis suppurativa, may also lead to the development of
keloids (Nemeth, 1993;Murray, 1994;English and Shenefelt,
1999;Bran et al., 2009). Isotretinoin is often used to treat acne
and has been suggested to act as an additional predisposing factor,
though this has not yet been proven conclusively (Guadanhim
et al., 2016). While burns have often been mentioned as one
of the many potential keloid-inducing events (Trusler and
Bauer, 1948;Nemeth, 1993;Murray, 1994;English and Shenefelt,
1999;Bran et al., 2009), they are usually associated with
the formation of widespread hypertrophic scars (Middelkoop
et al., 2011;Gold et al., 2014) rather than keloids. Fortunately,
venipuncture has not been reported to induce keloid scarring
(Yadav et al., 1995;Robles and Berg, 2007). Regardless of the
type of injury however, the resultant keloid scar response is
characteristically disproportionate to the original inciting injury
(Tuan and Nichter, 1998).
To summarize, keloid formation is most likely to occur after
an inciting stimulus such as dermal injury or inflammatory
process (environmental factor) at a keloid-prone anatomical
site (topological factor) in a genetically predisposed individual
(patient-related factor). The specific cellular abnormalities this
generates to ultimately result in keloid scar formation, are
discussed next (Figure 2).
Keloid Cellular Abnormalities
Although the keloid fibroblast is still considered the main culprit
responsible for keloid scar formation, recent studies have shifted
the focus to recognize the potential role of the epidermal
compartment and the immune system in keloidogenesis. This
section will address the trends in reported cellular abnormalities
across the entire spectrum of cells present in skin and/or involved
in wound healing (see Supplementary Tables S3A–D).
Abnormalities in Keloid Epidermal Cell Population
Often overlooked or even designated as ‘normal appearing’
(Butler et al., 2008b;Hollywood et al., 2010), the keloid
epidermis has only recently started garnering attention in
keloid research. A summary of the reported keloid epidermal
abnormalities has been listed in Supplementary Table S3A and
a summary of the histopathological epidermal abnormalities
can be found in the paragraph ‘Keloid Histopathology’ and
in Supplementary Table S1. The abnormalities of the keloid
epidermis are not limited to those visible by histopathology
alone, there is growing evidence that its barrier function is
also affected as measurements of transepidermal water loss
(TEWL) and high-frequency conductance suggest keloid scars
may show altered stratum corneum function compared with
healthy skin (Suetake et al., 1996;Sogabe et al., 2002). In
line with these findings, we found that the specific abnormal
overexpression of terminal differentiation marker involucrin was
not only associated with increased epidermal thickness, but
also with disorganization of the stratum corneum as visualized
by transmission electron microscopy (Limandjaja et al., 2017).
The following paragraphs will focus on abnormalities reported
for keloid keratinocytes and melanocytes. Langerhans cells also
reside in the epidermal compartment, but will be discussed in the
paragraph on keloid immune cells.
Keloid keratinocytes
Keloid keratinocytes may have a more direct role in keloid
scarring than previously assumed. Increased expression of
growth factors and cytokines such as CTGF (Khoo et al., 2006),
HGF and its receptor c-Met (Mukhopadhyay et al., 2010),
VEGF and PLGF (Ong et al., 2007b) have been demonstrated
in keloid-derived keratinocytes. Furthermore, cultured keloid
keratinocytes were found to differentially express 538 genes in
a study by Li and Wu (2016) and of these, further functional
analysis identified homeobox A7 (HOXA7), minichromosome
maintenance 8 (MCM8), proteasome subunit αtype 4 (PSMA4)
and proteasome subunit βtype 2 (PSMB2) as key differentially
expressed genes. In another gene expression study, Hahn
et al. (2013) found abnormal expression of genes involved in
differentiation, cell–cell adhesion and increased motility. Keloid
keratinocytes also contribute to keloid scarring by paracrine
regulation of ECM synthesis in fibroblasts, as evidenced by their
ability to induce a more profibrotic phenotype in vitro even in
fibroblasts of normal skin origin (Lim et al., 2002).
Lastly, several studies support a role for epithelial-
mesenchymal transition (EMT) in keloid scarring, a
phenomenon by which epithelial cells undergo phenotypic
changes and acquire more mesenchymal characteristics (Stone
et al., 2017). EMT has been found to occur in wound healing,
and plays a role in fibrosis by serving as a source of myofibroblast
generation (Stone et al., 2017). The changes associated with
EMT have been reported in keloid scars and involve the loss
of epithelial cell markers such as E-cadherin (Ma et al., 2015;
Yan et al., 2015;Hahn et al., 2016) and gain of mesenchymal
characteristics such as vimentin and FSP-1 (fibroblast specific
protein 1) expression (Yan et al., 2010, 2015;Ma et al., 2015;
Hahn et al., 2016;Kuwahara et al., 2016), combined with
changes in cell shape toward a more motile and migratory
phenotype (Hahn et al., 2013, 2016;Supp et al., 2014;Stone
et al., 2017). In short, the fundamental abnormalities found
in the keloid keratinocytes with respect to wound healing
mediator secretion, differentially expressed genes, paracrine
effects on co-cultured cells and epithelial-mesenchymal
transition, all support a more active role for keratinocytes
in keloidogenesis.
Keloid melanocytes
Little has been published on the role of melanocytes in
keloid pathogenesis (see Supplementary Tables S1, S3A),
despite the long observed increased keloid incidence in
individuals with darker pigmentation (Burd and Huang,
2005;Wolfram et al., 2009). To our knowledge, only Gao
et al. (2013) addressed the potential role of melanocytes
specifically in both hypertrophic and keloid scar formation
and proposed that during wound healing, a damaged
basement membrane allows the melanocytes to interact
with the dermal fibroblasts. The ensuing increase in
Frontiers in Cell and Developmental Biology | www.frontiersin.org 8May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 9
Limandjaja et al. The Keloid Disorder
fibroblast proliferation and collagen production together
with activation of the TGF-βpathway, promote abnormal scar
formation. They performed indirect co-culture experiments
in which melanocytes were able to induce increased levels
of proliferation, collagen I, TGF-β1 and its downstream
p-SMAD 2/3 expression in normal fibroblasts compared with
monocultured fibroblasts.
The increased melanin in keloid-prone patients may also
contribute to keloid scar formation by inhibiting the senescence-
inducing and anti-inflammatory effects of UVB radiation
(Wirohadidjojo et al., 2011) and vitamin D (Cooke et al.,
2005), respectively. However, variations in melanin levels
alone cannot fully explain the association of keloids with
darker pigmented individuals, as it has been reported that
the African albino population shows similar keloid prevalence
rates to the normally pigmented Africans (Kiprono et al.,
2015). Aberrations have also been reported in the steps
involved in melanogenesis, the process by which melanin is
generated (Koonin, 1964;Slominski et al., 1993;Song, 2014;
Wadhwa et al., 2016). For example, we now know that
polymorphisms in the MC1R gene are in fact responsible for
ethnic variations in pigmentation (Videira et al., 2013), and it
has been demonstrated that this receptor is not only expressed
on melanocytes but can also be found on dermal fibroblasts
(Stanisz et al., 2011). In fact, Luo et al. (2013) reported that
keloid scars and particularly keloid fibroblasts, showed reduced
expression of the melanocortin-1 receptor. They proposed that
this may negate the α-MSH-mediated suppression of collagen
synthesis and myofibroblast formation, thereby stimulating
keloid development. Additionally keloid fibroblasts have also
been found to be resistant to inhibitory effects of TGF-β1
on POMC expression (Teofoli et al., 1997;Lotti et al., 1999).
Thus, keloids association with increased pigmentation may very
well not reflect a primary abnormality in the melanocytes,
but a concomitant altered function of a shared receptor in
the fibroblasts.
Abnormalities in Keloid Dermal Cell Population
By their very nature, hypertrophic and keloid scars are defined
by the presence of raised, protruding scar tissue. The focus
of most studies has therefore understandably been on the
dermal component and more specifically on the extracellular
matrix (ECM) and the ECM-producing fibroblasts. Both keloid
and hypertrophic scars showed increased cellularity and were
in excess of all three primary ECM components of water,
collagen and proteoglycans. Notably, in keloids these processes
were significantly upregulated compared with normal skin and
hypertrophic scars (Ueda et al., 1999;Miller et al., 2003;
Meenakshi et al., 2005).
A summary of the reported keloid dermal abnormalities
has been listed in Supplementary Tables S3B–D, and a
summary of the dermal histopathological abnormalities
can be found in the paragraph ‘Keloid Histopathology’
and in Supplementary Table S1. The following paragraphs
will detail abnormalities reported for keloid fibroblasts and
keratinocyte/fibroblast interactions, myofibroblasts, fibrocytes,
endothelial cells, and nerve cells.
Keloid fibroblasts
An overwhelming number of studies have been devoted to
the keloid-derived fibroblast. However, the sheer multitude of
papers published on keloid fibroblasts have made it impossible
to discuss them all in this review and is outside our scope.
For this reason, the focus of this review was limited to
fibroblast abnormalities as they pertain to the main themes listed
by Marneros and Krieg (2004) and Robles and Berg (2007):
proliferation, ECM synthesis and degradation, expression of
wound healing mediators and apoptosis. A summary of these
publications is listed in Supplementary Table S3B, overall trends
in these in vitro monoculture findings will be summarized in the
following paragraphs.
There is an overall increase in the number of fibroblasts in
keloids, most likely mediated by increased proliferation rates.
Although some have also reported normal or even decreased
proliferation rates in keloid fibroblasts, the overwhelming
majority of in vitro monolayer studies support keloid fibroblast
hyperproliferation (Russell et al., 1988;Concannon et al., 1993;
Blume-Peytavi et al., 1997;Carroll et al., 2002;Carroll and
Koch, 2003;Giugliano et al., 2003;Hanasono et al., 2003,
2004;Meenakshi et al., 2005;Lim et al., 2006, 2009;Yeh
et al., 2006;Ghazizadeh et al., 2007;Ong et al., 2007a;Akino
et al., 2008;Witt et al., 2008;Zhang G. et al., 2009;Jing
et al., 2010;Romero-Valdovinos et al., 2011;He et al., 2012;
Syed and Bayat, 2012;Jurzak and Adamczyk, 2013;Wang
et al., 2013;Xin et al., 2017). In conjunction with increased
proliferation, reduced apoptosis by any means would also lead
to a cumulative net increase in the keloid fibroblast population.
Despite some papers reporting increased apoptosis (Akasaka
et al., 2000, 2005), overall, the majority of studies report
reduced apoptosis (Ladin et al., 1998;Messadi et al., 1999;
Chipev et al., 2000;Luo et al., 2001;Funayama et al., 2003;
Tucci-Viegas et al., 2010;Wang et al., 2013). Apoptosis is
also reduced in keloid fibroblasts by upregulation of apoptosis-
resistance (Ohtsuru et al., 2000;Messadi et al., 2004;Lu et al.,
2007b;Seifert et al., 2008), as well as telomere dysfunction
and defective senescence. Findings of telomerase upregulation
(Yu et al., 2016) and consequent telomere lengthening (Granick
et al., 2011;Yu et al., 2016) in keloid fibroblasts support
lifespan-prolonging effects of telomere dysfunction, although
telomere shortening as a result of oxidative stress has also
been reported in 30% of the keloids studied by De Felice
et al. (2009). In normal wound healing, fibroblasts eventually
become senescent and can then act as inhibitors in the regulation
of fibroblast proliferation and ECM synthesis. In this way,
defective senescence may also result in a net increase in fibroblast
density (Blaži´
c and Brajac, 2006), but literature on senescence
in keloid fibroblasts has been sparse and even counterintuitive
(Varmeh et al., 2011).
In line with their invasive nature, keloid fibroblasts also
show increased migration (Fujiwara et al., 2005a;Lim et al.,
2006;Witt et al., 2008;Wen et al., 2011;Syed and Bayat, 2012;
Wang et al., 2013;Fang et al., 2016;Jumper et al., 2017;Hsu
et al., 2018) and capacity for invasion in 3D invasion assays
(Dienus et al., 2010;He et al., 2012;Syed and Bayat, 2012;
Wang et al., 2018). Furthermore, increased metabolic activity
Frontiers in Cell and Developmental Biology | www.frontiersin.org 9May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 10
Limandjaja et al. The Keloid Disorder
(Meenakshi et al., 2005;Vincent et al., 2008), increased ECM
synthesis (McCoy et al., 1982;Abergel et al., 1987;Ala-Kokko
et al., 1987;Babu et al., 1989;Berman and Duncan, 1989;Suzawa
et al., 1992;Fujiwara et al., 2005a;Ong et al., 2007a;He et al.,
2012) and deposition (Abergel et al., 1985;Uzawa et al., 2003;
Fang et al., 2016) combined with reduced ECM degradation
(Abergel et al., 1985;Berman and Duncan, 1989;Uchida et al.,
2003;Yeh et al., 2006, 2009;Seifert et al., 2008;Russell et al., 2010;
McFarland et al., 2011;Suarez et al., 2013), all contribute to the
ECM overexpression and the resulting dermal protuberance that
defines these scars (see Supplementary Table S3B). Increased
levels of collagen I (Uitto et al., 1985;Ala-Kokko et al., 1987;
Lee et al., 1991;Friedman et al., 1993;Sato et al., 1998;Chipev
et al., 2000;Daian et al., 2003;Hasegawa et al., 2003;Lim et al.,
2003;Hsu et al., 2006, 2018;Xia et al., 2007;Zhang G. et al., 2009;
Dienus et al., 2010;McFarland et al., 2011;Lin et al., 2015;Suarez
et al., 2015;Fang et al., 2016;Luo et al., 2017), a major constituent
of the dermis, is likely responsible for the bulk of this increased
tissue mass. Additionally, there is an increased collagen I:III ratio
(Uitto et al., 1985;Abergel et al., 1987;Lee et al., 1991;Friedman
et al., 1993;Zhang G. et al., 2009), despite variable reports on
the levels of collagen III (Clore et al., 1979;Uitto et al., 1985;
Ala-Kokko et al., 1987;Lee et al., 1991;Sato et al., 1998;Lim
et al., 2002, 2003, 2013;Zhang G. et al., 2009;Lin et al., 2015;
Hsu et al., 2018). The most reported trend is that of increased
collagen III levels (Ala-Kokko et al., 1987;Lee et al., 1991;Sato
et al., 1998;Lim et al., 2003, 2013;Zhang G. et al., 2009;Hsu et al.,
2018). Other ECM constituents also expressed at higher levels in
keloid fibroblasts include fibronectin (Kischer and Hendrix, 1983;
Babu et al., 1989;Kischer and Pindur, 1990;Sible et al., 1994;
Blume-Peytavi et al., 1997;Chipev and Simon, 2002;Liang et al.,
2013;Suarez et al., 2015;Fang et al., 2016;Luo et al., 2017;Hsu
et al., 2018), elastin (Russell et al., 1989, 1995;Lee et al., 1991),
glycosaminoglycans (Berman and Duncan, 1989;Suarez et al.,
2013), and both small and large proteoglycans (Yagi et al., 2013).
The profibrotic characteristics of keloid fibroblasts are at
least in part, mediated by increased levels of several key
wound healing mediators and their associated receptors (see
Supplementary Table S3B). Major pathways upregulated in
keloid fibroblasts include TGF-β1 (Mccormack et al., 2001;
Mikulec et al., 2001;Carroll et al., 2002;Carroll and Koch, 2003;
Funayama et al., 2003;Hanasono et al., 2003;Xia et al., 2004;
Bock et al., 2005;Fujiwara et al., 2005b;Bran et al., 2010;Lim
et al., 2013;Wang et al., 2013;Jurzak et al., 2014;Yang et al.,
2014;Lin et al., 2015;Suarez et al., 2015;Fang et al., 2016), TGF-
β2 (Xia et al., 2005;Bran et al., 2010;Suarez et al., 2015;Fang
et al., 2016), and their receptors (Chin et al., 2001;Xia et al., 2004;
Tsujita-Kyutoku et al., 2005); CTGF (Xia et al., 2004, 2007;Khoo
et al., 2006;Russell et al., 2010;Jurzak et al., 2014;Fang et al.,
2016); VEGF (Wu et al., 2004;Fujiwara et al., 2005b;Ong et al.,
2007b;Dienus et al., 2010); interleukins IL-6 (Xue et al., 2000;
Ghazizadeh et al., 2007;Do et al., 2012) and IL-8 (Do et al., 2012);
as well as IGF-1 receptor (Yoshimoto et al., 1999;Ohtsuru et al.,
2000;Phan et al., 2003;Hu et al., 2014) and its binding-related
proteins (Phan et al., 2003;Seifert et al., 2008;Smith et al., 2008;
Russell et al., 2010). Moreover, keloid fibroblasts not only produce
higher levels of wound healing factors, they are also inherently
more sensitive to the effects of many of these factors (see
Supplementary Table S3B). Keloid fibroblasts showed increased
collagen secretion, PAI-1 and PDGFαreceptor expression, as well
as increased proliferation and migration in response to IL-18 (Do
et al., 2012), VEGF (Wu et al., 2004), TGF-β1 (Messadi et al.,
1998), HDGF (Ooi et al., 2010), and CTGF (Luo et al., 2017),
respectively, which was absent in normal fibroblasts. Similarly,
keloid fibroblasts exhibited a greater response in ECM synthesis,
proliferation, migration, invasion and inflammatory mediator
secretion to TGF-β(Bettinger et al., 1996;Daian et al., 2003), HGF
(Jin, 2014), PDGF (Haisa et al., 1994), and IL-18 (Do et al., 2012)
stimulation, respectively, compared with normal skin fibroblasts.
In a similar fashion to normal fibroblasts (Kroeze et al.,
2009), keloid-derived fibroblasts display mesenchymal stem cell
(MCS) markers and possess the multipotency to differentiate
into adipocytes, osteocytes, chondrocytes, smooth muscle cells,
endothelial cells, and neural lineage cells (Moon et al., 2008;
Iqbal et al., 2012;Plikus et al., 2017); thereby earning
the descriptor of multipotent precursor cells. Interestingly,
multipotency capabilities may differ between different scar
types as demonstrated by the ability of keloid fibroblasts,
but not their hypertrophic counterparts, to differentiate into
adipocytes either by stimulation with BMP4 or when co-
cultured with human scalp hair follicle cells (Plikus et al.,
2017). Iqbal et al. (2010) further differentiated between MSCs
of hematopoietic and non-hematopoietic origin, with the
majority comprising the non-hematopoietic subtype located
in the top and middle areas of the keloids. Regardless
of the MSC subtype, however, all MSC markers showed
progressive downregulation in culture with increasing cell
passage. Based on the similarly progressive loss of the keloid
phenotype with in vitro serial culturing and the abnormally
proliferative nature of keloid fibroblasts, Moon et al. (2008)
hypothesized that keloid fibroblasts may be stimulated by the
aberrant keloid cytokine milieu to remain in an undifferentiated
multipotent and proliferative stem cell state. By extension,
Qu et al. (2013) proposed that these keloid stem cells are
able to sustain themselves by asymmetric cell division due
to their drug resistant and high self-renewing abilities. The
continued generation of new aberrant keloid cells then sets
the typical tumor-like keloid growth in motion, and also
helps explain the high post-therapy recurrence rates. In
fact, the pathological keloid microenvironment may also be
responsible for generating the keloid stem cells in the first
place. Qu et al. (2013) also hypothesized that a pathological
niche exists in keloids that is the result of the pre-existing
abnormalities in keloid-prone patients, namely the enhanced
and persistent inflammatory response and the overexpression
of growth factors and their receptors. The multipotent keloid
fibroblasts, or rather keloid stem cells, are then transformed
from normal dermal stem cells after exposure to this pathological
keloid niche. Akino et al. (2008) co-culture experiments of
mesenchymal stem cells with keloid fibroblasts may support this
niche hypothesis, as mesenchymal stem cells showed similar
fibrotic and myofibroblast-like changes after exposure to keloid
fibroblasts in co-culture. Regardless of their cell of origin
however, the multipotent stem cell nature of the keloid fibroblast
Frontiers in Cell and Developmental Biology | www.frontiersin.org 10 May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 11
Limandjaja et al. The Keloid Disorder
appears to play an important role in the genesis and maintenance
of keloid scars.
Although highly informative, findings from fibroblast
monolayer cultures are not without their limitations. Serial
culturing, methods of fibroblast cell isolation (enzymatic vs.
explant), presence or absence of serum in culture medium,
and 3D vs. monolayer culturing, are all potential culturing
artifacts which have differed across studies and may influence
outcome parameters significantly. It is important to consider
the potential confounding effects of these culturing differences
while interpreting different results. As a final consideration, it
should be noted that keloid fibroblasts are usually compared with
fibroblasts derived from healthy non-lesional skin, while in fact
normotrophic scars represent the true standard against which
keloids should be compared. This holds true for all the tissue and
cellular components studied in keloid research, and is not limited
to the comparison of keloid fibroblasts to normotrophic scar-
derived fibroblasts. In conclusion, keloid fibroblast monocultures
have generated a multitude of interesting findings, but it is
important to remember the inherent limitations associated
with monoculture model systems and consider the influence of
differences in cell isolation and culture methods.
Abnormal keloid keratinocyte-fibroblast interactions
We know that the interactions between keratinocyte and
fibroblasts are an integral component of the normal wound
healing process (Lorenz and Longaker, 2003;Broughton et al.,
2006) and findings from in vitro double chamber co-culture
experiments have been particularly informative in this regard, as
they allow us to study indirect paracrine interactions between
the two cell populations (see Supplementary Table S4D).
Co-cultures of keratinocytes with fibroblasts show increased
proliferation, levels of ECM and growth factor expression
compared with monocultures (Phan et al., 2002, 2003;Funayama
et al., 2003;Lim et al., 2003;Xia et al., 2004;Khoo et al., 2006;
Ong et al., 2007b;Ooi et al., 2010;Do et al., 2012), but this effect is
generally greatest with keloid-derived cells. Keloid keratinocytes
are able to induce the profibrotic keloid phenotype in normal skin
fibroblasts (Lim et al., 2001, 2002, 2003;Funayama et al., 2003;
Xia et al., 2004;Khoo et al., 2006;Ong et al., 2007b;Chua et al.,
2011), while keloid fibroblasts are able to the propagate fibrosis
markers even when co-cultured with normal keratinocytes (Lim
et al., 2001;Phan et al., 2003;Xia et al., 2004;Supp et al., 2012b;
Lee Y.-S. et al., 2015).
These findings all strongly support a role for abnormal keloid
keratinocyte/fibroblast interactions in the pathogenesis of keloids
and thereby provide a new point of interception for therapeutic
strategies. In this light, Burd and Chan (2002) described an
interesting case of a pediatric patient with a giant keloid covering
most of the upper right leg and buttocks. In a multistep
procedure, the keloid tissue was removed and an artificial dermal
matrix was placed on the wound bed. This was followed by the
addition of an autologous keratinocyte cell suspension fixated by
a fibrin glue spray. During the 18-month follow-up there was
no recurrence of keloid formation. This case report serves as
an excellent example of a bench to bedside approach to negate
the adverse keloid epidermis-dermis interaction, by removing the
diseased dermal matrix and introducing normal keratinocytes.
Keloid myofibroblasts
Although Ehrlich et al. (1994) put forward the absence of
myofibroblasts as a feature that differentiates keloids from
hypertrophic scars, the opposite has also been observed (Santucci
et al., 2001). In fact, the overwhelming majority of studies report
the presence of α-SMA+myofibroblasts in 33–81% of the keloids
analyzed (Santucci et al., 2001;Kamath et al., 2002;Amadeu
et al., 2003;Lee J. Y. Y.et al., 2004;Moshref and Mufti, 2009).
Particularly when cultured in vitro, keloid fibroblasts can be
shown to contain a significant portion of myofibroblasts (Chipev
et al., 2000;Chipev and Simon, 2002;van der Slot et al., 2004;
Ong et al., 2007a;Lee et al., 2012;Jin, 2014;Suarez et al., 2015;
Luo et al., 2017;Shang et al., 2018). In our histopathological
study on abnormal scar types and immature scars (3–5 weeks
old), we identified a CD34/α-SMA+specific dermal cell
population, which were largely senescent in the abnormal scars
(p16+) but actively proliferating in the young scars (Ki67+)
(Limandjaja et al., 2019). See Supplementary Tables S2, S3B
for a summary of the histopathological results and cellular
abnormalities, respectively.
In wound healing by secondary intention, macrophages
stimulate wound bed-derived fibroblasts with TGF-β1 and PDGF
to transform them into myofibroblasts (Broughton et al., 2006).
In a recently published review, Lim et al. (2019) suggested
that the aberrant fibroblasts and myofibroblasts in keloids
may originate from an altogether different cell type, namely
the embryonal stem cell-like cell population located in the
endothelium of microvessels and on the perivascular cells
within keloid-associated lymphoid tissue. After injury, these
cells are thought to differentiate into abnormal fibroblasts and
myofibroblasts through the process of endothelial-mesenchymal
transition. Additionally, circulating fibrocytes or mesenchymal
stem cells from the bone marrow may also migrate to the
target site to generate the abnormal (myo-)fibroblast population.
In other words, myofibroblasts in the keloid environment
may have several sources of origin beyond the wound bed
fibroblasts. Alternatively, mesenchymal stem cells may also serve
as a myofibroblast source. Monolayer co-culture experiments
with mesenchymal stem cells and keloid fibroblasts have
shown that the latter are able to induce a myofibroblast-
like phenotypic switch of the mesenchymal stem cells (Akino
et al., 2008). In short, myofibroblasts may very well originate
from several different sources in addition to the wound-
bed fibroblasts.
In normal wound healing processes, we know that
myofibroblasts can produce significant wound surface reduction
through wound contraction (Lorenz and Longaker, 2003), but
Plikus et al. (2017) published an interesting new theory on how
myofibroblasts may contribute to the development of keloid
scars. They found that hair follicles are essential for inducing
myofibroblast-to-adipocyte reprogramming that allows for
regeneration rather than scar formation. As hair follicles are
absent from the keloid microenvironment, the myofibroblasts are
left unable to convert to adipocytes, thereby triggering the scar
response leading to keloid formation. In this way, hair follicles
and adipocytes may be involved in keloid scarring by effecting
myofibroblast dissipation from the wound bed, and serve as
interesting new potential therapeutic targets for further research.
Frontiers in Cell and Developmental Biology | www.frontiersin.org 11 May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 12
Limandjaja et al. The Keloid Disorder
Keloid fibrocytes
Bucala et al. (1994) were the first to suggest that the surrounding
connective tissue may not be the sole source of new fibroblasts
in wound repair and described a blood-borne cell with
fibroblast-like properties that enter sites of tissue repair (see
Supplementary Tables S1, S3B). These so-called fibrocytes were
characterized by a collagen+/vimentin+/CD34+phenotype and
not only produced ECM proteins and wound healing mediators,
but were also capable of acting as antigen-presenting cells and
differentiating into myofibroblasts (Bucala et al., 1994;Quan
et al., 2004). Based on the limited available literature, there are
increased numbers of CD45RO+/35F9+/MRP8/9+fibrocytes
in keloids compared with normotrophic scars (Iqbal et al.,
2012), and moreover, PBMCs derived from keloid patients
produced more LSP-1+/collagen1+fibrocytes than PBMCs
from healthy controls (Naylor et al., 2012). In further support
of this, Mathangi Ramakrishnan et al. (2012) showed that
keloid fibroblasts expressed increased levels of fibrocyte markers
(CD34+/CD86+), which were absent in normal fibroblasts.
This suggests an at least partial fibrocyte origin of the keloid-
derived fibroblasts. Reduced fibrocyte numbers have also been
reported (Ueda et al., 1999), but this was based on the
presence of histologically identified slender nuclei rather than
immunohistochemical phenotyping, and may therefore not be as
reliable. Given the aforementioned findings of increased fibrocyte
presence in keloids (Iqbal et al., 2012;Mathangi Ramakrishnan
et al., 2012) and their potential differentiation into abnormal
keloid myofibroblasts (Lim et al., 2019), fibrocytes may be
significantly involved in keloidogenesis and therefore deserve
further investigation.
Keloid endothelial cells
Both increased (Ehrlich et al., 1994;Amadeu et al., 2003;Tanaka
et al., 2004;Materazzi et al., 2007;Ammendola et al., 2013;
Bakry et al., 2014) and decreased (Beer et al., 1998;Ueda et al.,
2004;Bux and Madaree, 2010;Kurokawa et al., 2010) vasculature
have been reported in keloid scars in approximately equal
measure (see Supplementary Tables S2, S3C). However, based
on reports of microvessel occlusion and increased expression
of hypoxia-induced factor 1α(HIF-1α) in abnormal scars, it
has been proposed that keloids are relatively hypoxic tissues
(Zhao et al., 2017). The ischemia hypothesis builds on this to
explain how hypoxia can contribute to keloid scar development.
Kischer et al. (1982) demonstrated that unlike normal skin, the
overwhelming majority of hypertrophic and keloid scars have
microvessels with occluded lumens and that this was likely due
to endothelial cell proliferation. The authors considered this
reaffirmation of their hypothesis that hypoxia is an integral
factor in the generation of hypertrophic scar and keloid (Kischer
et al., 1982), but whether or not this relative hypoxia promotes
fibroblast and endothelial cell proliferation has still not been
determined (Song, 2014).
Kischer (1984) also suggested another mechanism by which
the microvessel abnormalities could generate both hypertrophic
scars and keloids. They proposed that injury leads to regeneration
of the microvessels, and suggest that the pericytes of the
newly regenerating microvessels form the source of the
fibroblasts generating the excessive collagen which characterizes
these abnormal scars.
It has also been suggested that endothelial cell dysfunction
plays a role in keloidogenesis. Ogawa and Akaishi (2016)
proposed that local factors such as stretching tension together
with genetic factors both act to induce endothelial cell
dysfunction in the form of vascular hyperpermeability during
the inflammatory phase of wound healing. This prolongs
the influx of inflammatory cells and factors, thereby also
prolonging the inflammatory phase. Consequently, dysfunction
of the fibroblast cell population leads to the development of
either hypertrophic or keloid scars. Lastly, endothelial cells
may also contribute to keloid scar development by undergoing
endothelial-mesenchymal transition (EndoMT) to acquire a
mesenchymal phenotype (Lee Y.-S. et al., 2015). In this way,
endothelial cells may directly serve as a source of the abnormal
keloid fibroblasts.
Keloid nerve cells
Based on the symptoms of itching and pain, both sensations
carried by small nerve fibers, there does appear to be a role
for nerve cells in the development of keloid scars. Yet, thus
far little has been published on the presence of nerve cells
in keloid tissue (see Supplementary Tables S2, S3C). Sensory
nerve fibers have also been mentioned in the context of Ogawas
mechanobiology theory on keloid pathogenesis (Ogawa, 2011).
As part of the group of skin receptors perceiving mechanical
forces, information from the sensory fibers is then relayed to the
central nervous system leading to the release of neuropeptides,
which can then modulate scarring by altering skin and immune
cell functions. However, studies staining for nerve fibers in
keloid tissue have reported both increased (Hochman et al., 2008;
Drummond et al., 2017) and decreased (Saffari et al., 2018) nerve
fiber densities. As different markers (PGP9.5 and S100 protein,
respectively) were used to identify the nerve fibers, this could in
part explain the different outcomes.
Keloid Immune Cells
Although both increased and reduced levels of
certain immune cell types have been reported (see
Supplementary Tables S2, S3D), overall there appears to
be an increase in macrophages (Boyce et al., 2001;Shaker et al.,
2011;Bagabir et al., 2012a;Jiao et al., 2015), T-lymphocytes
(Boyce et al., 2001;Shaker et al., 2011;Bagabir et al., 2012a;Jiao
et al., 2015) and mast cells (Kamath et al., 2002;Moshref and
Mufti, 2009;Shaker et al., 2011;Bagabir et al., 2012a;Ammendola
et al., 2013) in keloids that have been found to interact with
each other, other immune cells and dermal fibroblasts on a
cellular level (Boyce et al., 2001;Santucci et al., 2001;Shaker
et al., 2011). Moreover, macrophages and T-lymphocytes from
keloids also showed intrinsic abnormalities compared with their
normal counterparts. Keloid-derived macrophages showed a
high activation status, increased M2 polarization and overall
increased expression of both M1 and M2 activation factors
compared with normal skin macrophages (Jin et al., 2018).
They were also more potent at inducing the regulatory T-cell
phenotype when co-cultured with CD4+T-lymphocytes from
Frontiers in Cell and Developmental Biology | www.frontiersin.org 12 May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 13
Limandjaja et al. The Keloid Disorder
keloid patients’ blood. Thus, there was not just a general increase
in T-lymphocytes, but specifically the regulatory (Jin et al.,
2018) and memory (Chen et al., 2018) T-cells, as well as an
increased CD4+:CD8+ratio (Boyce et al., 2001;Bagabir et al.,
2012a) in keloids. Furthermore, the altered cytokine production
in keloid-derived memory T-cells (Chen et al., 2018) and the
reduced mitogenic response of circulating T-cells to known
mitogenic stimuli (Bloch et al., 1984), suggests that an abnormal
T-cell response may contribute to keloid scarring. In this way,
the previously discussed sebum reaction hypothesis is also
an extension of this concept as the intrinsically abnormal,
sebum-sensitive T-lymphocytes take center stage in this theory
(Song, 2014).
Mast cells have also been found in abundance in keloid scars.
Arbi et al. (2015) found that mast cells were closely associated
with fibroblasts in keloid scars and that the phagocytosis of
collagen fibrils by mast cells was a common ultrastructural
feature. They hypothesized that the abnormal collagen synthesis
observed in keloids and the consequent accumulation of collagen
fibers, are able to induce increased mast cell recruitment
and subsequent collagen phagocytosis. The resulting release
of mast cell-derived mediators (interleukins, mediators, and
growth factors) is then able to stimulate further collagen
production and thereby aids further keloid scar development.
Lastly, there have been variable reports on the presence of
Langerhans cells in the epidermal compartment of keloid
scars, both normal (Bagabir et al., 2012a) and increased (Jiao
et al., 2015) numbers have been observed in both hypertrophic
and keloid scars.
Several hypotheses centering on inflammatory processes
have also been put forward. In the chronic inflammation
hypothesis, Dong et al. (2013) posed that the presence of
chronic inflammation in keloids indicates that local inflammation
promotes keloid formation. The traumatic and inflammatory
stimuli that trigger keloid scar formation result in the
continuous upregulation of already highly sensitive pro-
inflammatory genes in the keloid microenvironment. This keloid
microenvironment fosters the development of abnormalities
in the resident keloid fibroblast, which in turn is considered
to be the driving force behind keloid scar formation. Ogawa
(2017) has further expanded on this notion and postulated
that chronic inflammation is responsible for the invasive
growth of keloids and even suggests that both hypertrophic
scars and keloids are principally inflammatory disorders of
the reticular dermis rather than being skin tumors. In
the neurogenic inflammation hypothesis, the inflammation
is thought to arise from mechanical stress, such as skin
stretching, which stimulates mechanosensitive receptors on
sensory fibers to release neuropeptides. These then bind to
receptors of various skin cell types including keratinocytes
and fibroblasts, mast cells and endothelial cells. Vasodilation
and vessel permeabilization, increased mast cell release of
histamines and cytokine production (including TGF-β) take
place as a result of this. Fibroblasts then become activated as
a result of the neurogenic inflammation and the upregulation
TGF-β, leading to keloid and hypertrophic scar formation
(Akaishi et al., 2008).
Other Proposed Hypotheses on Keloid
Scar Formation
The myriad of available treatment modalities is only matched
by the multitude of proposed hypotheses to explain keloid
scar formation. These are not mutually exclusive, and further
support the notion that keloid scarring has a multifactorial
genesis (Slemp and Kirschner, 2006). When appropriate, these
have already been discussed in the appropriate paragraphs of
the section on keloid cellular abnormalities. A brief discussion
of additional hypotheses that could not be categorized in the
previous paragraph, will follow next.
Keloid Triad Hypothesis
Perhaps one of the only proposed hypotheses encompassing
multiple risk factors in one, the keloid triad hypothesis
(Agbenorku et al., 1995) is defined as a group of three etiologic
factors: genetic links, infective agent (bacterial, viral, or other)
and surgery (e.g., sutures, tension of suture lines, location of
sutures in relation to the relaxed skin tension lines); which must
be simultaneously present and interact to develop keloid scarring.
These three factors are further subdivided into major factors
and minor etiological factors. Major factors include: African
ethnicity, age 10–30 years, familial susceptibility or keloid-prone
upper part of body site. Minor factors include: orientation
of incisions/sutures with respect to RTSLs, wound or sutures
under tension, healing by secondary intention, type of infection;
determines whether or not a keloid scar is likely to develop. At
least one major and two minor factors must be present for keloid
scars to develop. A keloid is unlikely to develop if all three factors
are minor or if only two factors are present, but a hypertrophic
scar may form instead.
Incomplete Malignancy Hypothesis
When Ladin et al. (1998) studied apoptosis in keloid scars, they
found that the level of apoptotic cells was significantly reduced
in keloid tissue and fibroblasts compared with normal foreskin
tissue and fibroblasts. However, keloid fibroblasts did show
increased apoptosis upregulation in response to treatment with
hypoxia, hydrocortisone or γinterferon, while normal fibroblasts
were only responsive to high doses of hydrocortisone. Because of
this, Ladin et al. (1998) suggested that keloids may represent a
type of incomplete malignancy that has undergone some, but not
all tumourigenic changes.
Viral Hypothesis
In this infection-based hypothesis (Alonso et al., 2008), the
authors proposed a role for a normally quiescent, unknown
virus in the bone marrow or lymphatic system which is
activated in a genetically susceptible person with a wound.
This virus can then reach the wound via fibrocytes that are
chemoattracted to the wound site or via infecting virions
in the saliva arriving at the wound bed. There the many
chemical stimuli from the wound healing processes allow the
virus to become activated, resulting in transcription of viral
proteins which derail wound healing and eventually lead to
keloid scarring.
Frontiers in Cell and Developmental Biology | www.frontiersin.org 13 May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 14
Limandjaja et al. The Keloid Disorder
Stiffness Gap Hypothesis
Born out of recent findings on the role of mechanotransduction
in keloid scarring, this theory (Huang et al., 2017) proposes that
the enlarged gap between ECM stiffness and cellular stiffness
enables the constant and continued keloid progression. The
ECM is not only a cellular scaffold storing important wound
healing mediators, but its rigidity also influences fibroblast
function and can induce processes such as fibroblast migration,
proliferation and differentiation. When dermal fibroblasts sense
the stiffness gaps between the ECM and themselves via
mechanotransduction, alterations in fibroblast phenotype ensue
which promote proliferation, migration and ECM synthesis and
therefore contribute to keloid progression. However, systematic
mechanobiological experiments to verify this hypothesis have yet
to be performed.
Immunonutritional Hypothesis
It has been proposed by several authors that nutritional
imbalances can promote prolonged inflammation and cytokine-
mediated reactions which contribute to keloid scarring (Huang
and Ogawa, 2013a). Nutritional imbalances have been reported
for essential fatty acids and micronutrients in keloid-prone
patients. For example, the keloid-prone black South African
population were deficient in the essential fatty acids a-linolenic
acid, eicosapentaenoic acid and docosahexaenoic acid, but
showed an increased intake of linoleic acid and arachidonic acid.
The levels of micronutrients (calcium, copper, iron, vitamin B2,
vitamin C, vitamin A, and zinc) was also insufficient in their diet
(Louw and Dannhauser, 2000), and this was also reflected in the
serum as well as keloid scar tissue (Bang et al., 2002). Compared
with non-keloid forming patients, keloid-prone patients showed
higher levels of zinc in their serum, higher copper levels in non-
lesional normal skin and in the keloid scar tissue. The essential
fatty acid imbalance has also been observed in the membrane
of keloid fibroblasts with lower linoleic acid and higher neutral
lipid levels in the central keloid area compared with its periphery
(Louw et al., 1997). Lipids are not only essential cell membrane
constituents, but also have important roles in local inflammation
and intracellular signal transduction. As such they are also active
participants in the chronic inflammatory processes that result
in the development and continued growth of keloids (Huang
and Ogawa, 2013b). And finally, dietary compounds have been
shown to affect keloid-derived cells at the in vitro level as well.
Green tea and its extracts have been shown to reduce keloid
fibroblast proliferation, migration and collagen synthesis (Zhang
et al., 2006;Park et al., 2008).
Metabolic Hypothesis
It has also been theorized that the net overexpression of
ECM components in the keloid dermis is also the result of
both the increased number of fibroblasts and their increased
intrinsic metabolic activity, aside from the usual suspects
in abnormal signaling pathways (Butler et al., 2008a). Both
keloids and hypertrophic scars showed increased numbers of
fibroblasts compared with normal scars or normal skin (Ueda
et al., 1999), but only keloid fibroblasts showed increased
metabolic activity. Metabolic activity has also been measured
as levels of ATP (Ueda et al., 1999;Vincent et al., 2008) and
as total protein content synthesis together with endoplasmic
reticulum staining (Meenakshi et al., 2005), all of which have
proven to be upregulated in keloids compared with normal
skin, normotrophic and hypertrophic scars. Modulating keloid
metabolism may therefore be a way to reduce keloid scar growth
(Butler et al., 2008a;Huang and Ogawa, 2013a).
Nitric Oxide Activity
Using mathematical modeling, Cobbold (2001) proposed that
nitric oxide could be involved in both hypertrophic and keloid
scar formation. Nitric oxide is a known source of free radical
molecules which stimulates collagen synthesis in normal wound
healing, it therefore follows that excessive amounts of nitric
oxide can result in the elevated scar tissue characteristic of these
abnormal scar types. Cobbold further speculated that the source
of this excessive nitric oxide may be the basal epidermis, in
part because the generated free radicals also stimulate melanin
synthesis and melanocytes are located within this epidermal layer.
Psychoneuroimmune-Endocrine Hypothesis
Hochman et al. (2015) posed that the “brain-skin connection”
may play a role in keloid scarring. The psychogenic component is
a part of the pathogenesis in most skin diseases such as psoriasis
and atopic dermatitis, where it can trigger integrated responses
from the nervous, immune and the endocrine systems. Normal
wound healing also depends on neurogenic factors, which in turn
are influenced by psychological, immune and endocrine factors.
Any change in these factors may interfere with the scar formation
process. Hochman speculated that stressed patients themselves
exacerbate the neuro-endocrine system, leading to the release of
hormones, neural transmitters, and immune cells which creates
an inflammatory microenvironment that stimulates fibrotic
processes (Hochman et al., 2015). This has not been extensively
studied for abnormal scars but, stress as measured by the sweating
response during a stress-inducing task was found to be associated
with an increase in keloid recurrence rates after surgery with
postoperative radiotherapy (Furtado et al., 2012).
Hypertension Hypothesis
Dustan (1995) first hypothesized that the differences in growth
factor abnormalities that are conducive to keloid scarring in the
black population, could also be responsible for the development
of hypertension and therefore explain the differences in
hypertension severity between the black and white population.
A significant increase in hypertension rates has in fact been
reported in keloid-prone patients compared with non-keloid
formers in both the African–American (Snyder et al., 1996;
Adotama et al., 2016;Rutherford and Glass, 2017) and Caucasian
population (Snyder et al., 1996). Ogawa et al. (2013) also reported
that severe keloid cases (multiple or large >10 cm2) in Japan were
significantly more likely to have hypertension than patients with
mild keloids (<2 or <10 cm2). Furthermore, in patients with
keloids, those with more severe hypertension were significantly
more likely to have multiple and/or larger keloids than patients
with normotension or less severe hypertension (Arima et al.,
2015). The mechanisms by which hypertension could lead to
Frontiers in Cell and Developmental Biology | www.frontiersin.org 14 May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 15
Limandjaja et al. The Keloid Disorder
FIGURE 3 | Keloid scar models. Overview of available keloid scar models. Keloid models can be further subdivided into in vivo and in vitro models, which can be of
human or animal origin. Human in vivo models include non-invasive testing methods (e.g., imaging and microscopy) and invasive methods, such as serial biopsy or
inducing keloid formation by wounding. Animal in vivo models include implantation of keloid tissue fragments, fibroblasts or full thickness skin equivalents, as well as
inducing keloid scar development by irritation or wounding. In vitro keloid models range from simple monolayers to 3D structures and co-culture systems. Indirect
co-culture systems include monolayer keloid fibroblasts combined with either monolayer keratinocytes or a fully differentiated epidermis. Lastly, explant cultures are a
combination of in vivo and in vitro as the method involves the maintaining of keloid tissue fragments in in vitro culture. Figure first published in Limandjaja (2019),
used with permission.
keloid development remain largely speculative for the time being,
but the changing blood flow likely affects the skin cellular
constituents in ways that promote fibrotic processes (Huang
and Ogawa, 2014). Angiotensin-converting enzyme (ACE) was
proposed as a potential common mechanism for the pathogenesis
of both keloids and hypertension, but Stewart and Glass (2018)
found no correlation between plasma ACE levels and keloids
or hypertension. Nevertheless, antihypertensives such as ACE-
inhibitors (e.g., captopril), calcium-channel blockers (e.g.,
verapamil) have been able to reduce keloid symptoms and are
part of the current therapeutic arsenal for keloid treatment.
Huang and Ogawa (2014) reviewed the evidence for the
Frontiers in Cell and Developmental Biology | www.frontiersin.org 15 May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 16
Limandjaja et al. The Keloid Disorder
participation of systemic hypertension in hypertrophic and
keloid scar pathogenesis and although they could not establish
a strong causal relationship, hypertension was considered a likely
risk factor for abnormal scarring.
MODELS OF KELOID SCARRING
Keloid scar models can largely be divided into in vivo and in vitro
models, Supplementary Tables S4A–E and Figure 3 give an
overview of the currently available keloid scar models.
In vivo Keloid Models
In vivo keloid models are of the human or animal
variety. Currently available in vivo human keloid models
(Supplementary Table S4A) can be further subdivided into
non-invasive, invasive, and computational models. Non-invasive
in vivo modeling generally comprises live imaging whereby
certain tissue characteristics are visualized. Invasive keloid
modeling varies from the relatively minor FDG injections to
measure glucose metabolism (Ozawa et al., 2006), to serial
biopsies to evaluate keloid development over time (Lin et al.,
2015) and even includes attempts to induce keloid formation in
known keloid formers by wounding (Mancini and Quaife, 1962).
Despite their value especially in the clinical setting for follow-up
evaluation, human in vivo models post-1962 are inherently
limited in the ability to manipulate experimental variables and
yield primarily observational data. While Lebeko et al. (2019)
saw particular merit in the temporal data acquired from serial
keloid biopsy analysis as an in vivo ‘4D model’ in their review on
keloid models, it is important to recognize the potential risk of
exacerbating the existing keloid scars by continual provocation
with biopsy-associated injury.
In animal models, keloid formation is either induced
or involves the implantation of human keloid tissue (see
Supplementary Table S4C). Unfortunately, inducing keloid scar
formation has proven virtually impossible (Seo et al., 2013)
and more often than not, hypertrophic scars developed instead
of keloids (Morris et al., 1997;Khorshid, 2005). Implanting
human keloid cells or tissue fragments into animal models
has been more successful (Hillmer and Macleod, 2002) and
resulted in the development of a palpable nodule-like mass.
While implanted keloid tissue was generally able to retain the
keloid-specific keloidal collagen even within the animal model,
the use of already established keloid tissue does not allow us
to study de novo keloid development. However, when tissue
culture techniques were combined with IL-6 exposure to implant
a keloid fibroblasts-hydrogel suspension into nude mice, the
resultant mass not only showed the greatest increase in growth,
but also gave rise to de novo keloidal collagen formation (Zhang
Q. et al., 2009). The appeal of an in vivo animal model is
obvious, but there is an important limitation to the use of any
animal for the study of keloid scarring, in that keloids occur
exclusively in humans (Tuan and Nichter, 1998;Yang et al.,
2003). Though exuberant granulation tissue on horse limbs has
often been posed as the equine version of keloid tissue, a recent
study (Theoret et al., 2013) comparing keloids and exuberant
granulation tissue has shown they are not in fact one and the
same. Fundamental differences in essential skin physiology (hair
follicle density, epidermal, and dermal thickness) and wound
healing mechanisms between rodents and humans (Perez and
Davis, 2008;Ramos et al., 2008) pose additional important
limitations. The in vivo microenvironment in animal models may
therefore provide a less than human-like exposure to the keloid
cell population.
Keloid Explant Models
Keloid tissue explants do not necessarily require implantation
into an animal model in order to survive and be used as a keloid
model in and of itself, they can be maintained in culture after
removal from the human body for up to 6 weeks (Duong et al.,
2006;Bagabir et al., 2012b) (see Supplementary Table S4B).
Various culture methods have been tried, but keloid morphology
appears best conserved in explants embedded in collagen gel
that are cultured air-exposed (Duong et al., 2006;Bagabir et al.,
2012b;Mendoza-Garcia et al., 2015). Unfortunately, relying on
fresh keloid tissue for this explant model system has its obvious
logistical issues with limited availability and potential inter-
and intralesional variability between samples. Another important
limitation of this model type is the absence of circulatory system,
as is the case with all in vitro models to date. Ultimately, this
model system appears best suited for the testing of therapeutics
(Syed et al., 2012, 2013a,b) rather than studying the pathogenetic
mechanisms underlying keloid formation, as there is very limited
ability to manipulate any experimental variables.
In vitro Co-culture Keloid Models
As skin physiology, immunology and wound healing is markedly
different in animals (Hillmer and Macleod, 2002;Ramos et al.,
2008;Seo et al., 2013;Seok et al., 2013) and since keloids
are exclusive to humans (McCauley et al., 1992;Yang et al.,
2003), there is a need for relevant human in vitro models (see
Supplementary Tables S4D, E). What is more, in vitro co-culture
systems allow for extensive experimental manipulation, such
as the development of heterotypic models combining normal
keratinocytes with keloid fibroblasts to study the individual
contribution of keloid fibroblasts to keloid formation.
Indirect co-cultures consist of double chamber systems
with keratinocytes seeded (as a monolayer or differentiated
epidermis) onto an upper porous transwell insert and fibroblasts
cultured in a monolayer on the underlying bottom well
(see Supplementary Table S4D). These indirect co-cultures
are simple-to-execute experiments that are particularly suited
to study paracrine interactions, as previously discussed in
Abnormal Keratinocyte-Fibroblast Interactions’. However, the
monolayer fibroblast cultures and the artificial separation
between keratinocytes and fibroblasts, bear little resemblance to
the true in vivo situation.
Direct co-cultures represent the most in vivo-like in vitro
culture model and include either mixed monolayer cultures
or full thickness skin equivalents comprised of keloid-derived
cells (see Supplementary Table S4E). The full thickness skin
equivalents comprised entirely of keloid-derived keratinocytes
and fibroblasts most closely resemble the native keloid, but thus
Frontiers in Cell and Developmental Biology | www.frontiersin.org 16 May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 17
Limandjaja et al. The Keloid Disorder
far, these have only been developed in the context of implantation
into an animal model (Supp et al., 2012b;Lee Y.-S. et al.,
2015) and/or subjected only to limited experimental analysis
(Supp et al., 2012a). Overexpression of ECM components,
particularly collagen I, was the common denominator in these
keloid models. Recently, we were also able to reconstruct keloids
in vitro with keloid-derived keratinocytes and fibroblasts, using
a collagen-elastin scaffold (Limandjaja et al., 2018a). Compared
with in vitro reconstructed normal skin, the keloid model showed
a trend of increased dermal thickness, increased α-SMA and p16
expression, reduced HGF secretion and reduced collagen type IV
α2 chain, hyaluronan synthase 1 and matrix metalloproteinase
3 gene expression. More importantly, the keloid model behaved
differently from similarly reconstructed hypertrophic scars, and
was able to demonstrate intralesional heterogeneity within
keloids (Limandjaja et al., 2018b). In other words, a relatively
simple in vitro keloid model was already able to demonstrate
certain intrinsic abnormalities in keloid-derived keratinocytes
and fibroblasts and in doing so, identified certain aspects of keloid
behavior which could not have been deduced from ex vivo biopsy
analysis alone. In our opinion, this both illustrates and validates
the use of skin tissue engineering as an important adjunct to
furthering keloid scar research.
Ultimately, there is no single universal keloid model
that is able to satisfactorily answer every experimental
objective (Hillmer and Macleod, 2002;Marttala et al., 2016).
Overall, animal models are better suited for therapeutic
testing and particularly for safety testing prior to human
administration (Hillmer and Macleod, 2002;Marttala et al.,
2016), while tissue culture systems are best suited to study
keloid pathogenesis (Hillmer and Macleod, 2002;Marttala
et al., 2016). Given the exclusive occurrence of keloids in
humans (Tuan and Nichter, 1998;Yang et al., 2003) and
other serious limitations of using animal models for wound
healing studies, as well as the fact that research on its
pathogenesis cannot rely solely on (immunohistochemical)
analysis of an intact native specimen; it seems pertinent
to focus our efforts on further advancing tissue culture
models - in particular the in vitro, organotypic full thickness
keloid skin equivalents. Once proven faithful to actual
in vivo keloid biology, using this model to study keloid
formation will eventually aid in the development of superior
treatment methods by uncovering new mechanisms of
pathogenesis. In addition, such a model would have the
added benefit of serving as a test object for future therapeutic
modalities without the complication of ethical objections
(Yang et al., 2003).
THE KELOID DISORDER
The International Classification of Disease (ICD) represents
the international standard for reporting diseases and health
conditions and in the most recently released ICD-11 (2018),
keloids have been grouped under ‘fibromatoses and keloids,
a heterogenous group of disorders characterized by increased
deposition of fibrous tissue in the skin and subcutaneous tissues.
This is departure from the previous ICD-10 (2016) in which
keloids were still categorized under ‘hypertrophic disorders of
skin, and is the result of recent international agreement to refer
to keloids as part of a ‘keloid disorder’ rather than a ‘keloid
scar.’ This change in ICD classification and nomenclature reflects
growing consensus among keloid researchers and clinicians alike
that keloids more closely resemble a benign, fibrous tumor
than scar tissue and can best be described in terms of a
fibrotic disorder.
By their very definition of continued, invasive horizontal
growth (Burd and Huang, 2005), keloids clearly share certain
characteristics of cancerous growths. Lu et al. (2007b) have
previously described keloids as behaving clinically as
non-metastatic malignancies, although histologically they
are benign. Similarly, Ladin et al. (1998) coined keloids
‘incomplete malignancies’ as keloid fibroblasts are still apoptosis-
responsive to in vitro treatment despite overall reduced
levels of apoptosis. The term refers to the manner in which
keloids show some, but not all the changes associated with
tumourigenesis. The absence of metastasis despite their invasive
growth and the aforementioned cancer-like characteristics,
separates keloids from true malignant tumors and lends
further support to the ‘incomplete malignancy’ term coined
by Ladin et al. (1998).
The change in nomenclature reflected in the newest edition
of the ICD may seem like an exercise in semantics, but it
should be noted that the cosmetic association with the word
‘keloid scar’ significantly complicates both insurance coverage
for treatment as well as funding for research. This is not to say
this nomenclature change must result in an altogether ban on
the association of the word ‘scar’ with keloids, merely that we
shift the focus away from a purely cosmetic association, toward
broadening our perspective of keloids for the fibroproliferative
disorders that they are and in this way recognize the severity of
this fibrotic disorder.
Keloid research remains plagued by significant controversy
and contradictory findings, which has been highlighted
throughout this manuscript where possible. Much of this could
be resolved by the standardization of study designs and general
research approach. Important considerations for future research
include (i) inclusion of a description of the keloid phenotype
(based on growth pattern), (ii) clarification of the location within
the keloid from which samples were taken (e.g., growing/bulging
area versus more quiescent region; peripheral versus central
areas), (iii) inclusion of scars of similar maturation stage (at
least 1 year old). Lastly, the inclusion of both hypertrophic
and normotrophic scars would be a valuable addition to keloid
research, as this provides a more comprehensive view of the
entire scarring spectrum.
AUTHOR CONTRIBUTIONS
GL, SG, and FN conceived the manuscript outline. GL performed
the literature study and wrote the manuscript with input from
all the authors. SG, FN, and RS supervised, helped shape the
manuscript, and offered final critical revision.
Frontiers in Cell and Developmental Biology | www.frontiersin.org 17 May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 18
Limandjaja et al. The Keloid Disorder
FUNDING
This study was financed by a grant from the Dutch
government: Rijksdienst voor Ondernemend Nederland, project
number INT102010.
SUPPLEMENTARY MATERIAL
The Supplementary Material for this article can be found online
at: https://www.frontiersin.org/articles/10.3389/fcell.2020.00360/
full#supplementary-material
REFERENCES
Abergel, R. P., Chu, M. L., Bauer, E. A., and Uitto, J. (1987). Regulation of
collagen gene expression in cutaneous diseases with dermal fibrosis: evidence
for pretranslational control. J. Invest. Dermatol. 88, 727–731. doi: 10.1111/
1523-1747.ep12470397
Abergel, R. P., Pizzurro, D., Meeker, C. A., Lask, G., Matsuoka, L., Minor, R. R., et al.
(1985). Biochemical composition of connective tissue in keloids and analysis
of collagen metabolism in keloid fibroblast cultures. J. Invest. Dermatol. 84,
384–390. doi: 10.1111/1523- 1747.ep12265471
Adotama, P., Rutherford, A., and Glass, I. I. D. A. (2016). Association of keloids
with systemic medical conditions: a retrospective analysis. Int. J. Dermatol. 55,
e37–e55. doi: 10.1111/ijd.12969
Agbenorku, P. T., Kus, H., and Myczkowski, T. (1995). The keloid triad hypothesis
(KTH): a basis for keloid etiopathogenesis and clues for prevention. Eur. J. Plast.
Surg. 18, 301–304.
Akaishi, S., Ogawa, R., and Hyakusoku, H. (2008). Keloid and hypertrophic scar:
neurogenic inflammation hypotheses. Med. Hypotheses 71, 32–38. doi: 10.1016/
j.mehy.2008.01.032
Akaishi, S., Ogawa, R., and Hyakusoku, H. (2010). Visual and pathologic analyses
of keloid growth patterns. Ann. Plast. Surg. 64, 80–82. doi: 10.1097/SAP.
0b013e31819967ed
Akasaka, Y., Fujita, K., Ishikawa, Y., Asuwa, N., Inuzuka, K., Ishihara, M.,
et al. (2001). Detection of apoptosis in keloids and a comparative study on
apoptosis between keloids, hypertrophic scars, normal healed flat scars, and
dermatofibroma. Wound Repair Regen. 9, 501–506. doi: 10.1046/j.1524-475x.
2001.00501.x
Akasaka, Y., Ishikawa, Y., Ichiro, O., Fujita, K., Masuda, T., Asuwa, N., et al.
(2000). Enhanced expression of caspase-3 in hypertrophic scars and keloid:
induction of caspase-3 and apoptosis in keloid fibroblasts in vitro. Lab. Investig.
80, 345–357. doi: 10.1038/labinvest.3780039
Akasaka, Y., Ito, K., Fujita, K., Komiyama, K., Ishikawa, Y., Akishima, Y.,
et al. (2005). Activated caspase expression and apoptosis increase in keloids:
cytochrome c release and caspase-9 activation during the apoptosis of keloid
fibroblast lines. Wound Repair Regen. 13, 373–382. doi: 10.1111/j.1067-1927.
2005.130404.x
Akino, K., Akita, S., Yakabe, A., Mineda, T., Hayashi, T., and Hirano, A. (2008).
Human mesenchymal stem cells may be involved in keloid pathogenesis. Int. J.
Dermatol. 47, 1112–1117. doi: 10.1111/j.1365-4632.2008.03380.x
Alaish, S. M., Yager, D. R., Diegelmann, R. F., and Cohen, I. K. (1995). Hyaluronic
acid metabolism in keloid fibroblasts. J. Pediatr. Surg. 30, 949–952. doi: 10.1016/
0022-3468(95)90319- 4
Ala-Kokko, L., Rintala, A., and Savolainen, E. R. (1987). Collagen gene expression
in keloids: analysis of collagen metabolism and type I, III, IV, and V procollagen
mRNAs in keloid tissue and keloid fibroblast cultures. J. Invest. Dermatol. 89,
238–244. doi: 10.1111/1523- 1747.ep12471056
Al-Attar, A., Mess, S., Thomassen, J. M., Kauffman, C. L., and Davison,S. P. (2006).
Keloid pathogenesis and treatment. Plast. Reconstr. Surg. 117, 286–300.
Ali, S. S., Hajrah, N. H., Ayuob, N. N., Moshref, S. S., and Abuzinadah, O. A. (2010).
Morphological and morphometric study of cultured fibroblast from treated and
untreated abnormal scar. Saudi Med. J. 31, 874–881.
Alibert, J. L. (1825). Description des Maladies de la Peau: Observées à
l’Hôpital Saint-Louis, et Exposition des Meilleures Méthodes Suivies Pour
Leur Traitement, Tome Second. Available online at: https://archive.org/details/
descriptiondesma02alib/page/100
Alonso, P. E., Rioja, L. F., and Pera, C. (2008). Keloids: a viral hypothesis. Med.
Hypotheses 70, 156–166.
Amadeu, T., Braune, A., Mandarim-de-Lacerda, C., Porto Cristovao, L. C.,
Desmoulière, A., and Costa, A. M. A. (2003). Vascularization pattern in
hypertrophic scars and keloids: a stereological analysis. Pathol. Res. Pract. 199,
469–473. doi: 10.1078/0344- 0338-00447
Ammendola, M., Zuccalà, V., Patruno, R., Russo, E., Luposella, M., Amorosi,
A., et al. (2013). Tryptase-positive mast cells and angiogenesis in keloids: a
new possible post-surgical target for prevention. Updates Surg. 65, 53–57. doi:
10.1007/s13304-012- 0183-y
Appleton, I., Brown, N. J., and Willoughby, D. A. (1996). Apoptosis, necrosis, and
proliferation: possible implications in the etiology of keloids. Am. J. Pathol. 149,
1441–1447.
Arbi, S., Eksteen, E. C., Oberholzer, H. M., Taute, H., and Bester, M. J. (2015).
Premature collagen fibril formation, fibroblast-mast cell interactions and mast
cell-mediated phagocytosis of collagen in keloids. Ultrastruct. Pathol. 39, 95–
103. doi: 10.3109/01913123.2014.981326
Arima, J., Huang, C., Rosner, B., Akaishi, S., and Ogawa, R. (2015). Hypertension:
a systemic key to understanding local keloid severity. Wound Repair Regen. 23,
213–221. doi: 10.1111/wrr.12277
Babu, M., Diegelmann, R., and Oliver, N. (1989). Fibronectin is overproduced by
keloid fibroblasts during abnormal wound healing. Mol. Cell. Biol. 9, 1642–
1650. doi: 10.1128/mcb.9.4.1642
Bagabir, R., Byers, R. J., Chaudhry, I. H., Müller, W., Paus, R., and Bayat,
A. (2012a). Site-specific immunophenotyping of keloid disease demonstrates
immune upregulation and the presence of lymphoid aggregates. Br. J. Dermatol.
167, 1053–1066. doi: 10.1111/j.1365-2133.2012.11190.x
Bagabir, R., Syed, F., Paus, R., and Bayat, A. (2012b). Long-term organ culture
of keloid disease tissue. Exp. Dermatol. Denmark 21, 376–381. doi: 10.1111/
j.1600-0625.2012.01476.x
Bakry, O. A., Samaka, R. M., Basha, M. A., Tharwat, A., and El Meadawy, I.
(2014). Hematopoietic stem cells: do they have a role in keloid pathogenesis?
Ultrastruct. Pathol. Engl. 38, 55–65. doi: 10.3109/01913123.2013.852646
Balci, D., Inandi, T., Dogramaci, C., and Celik, E. (2009). DLQI scores in patients
with keloids and hypertrophic scars: a prospective case control study. J. Dtsch.
Dermatol. Ges. 7, 688–692. doi: 10.1111/j.1610- 0387.2009.07034.x
Bang, R. L., Al-Bader, A. L., Sharma, P. N., Matapallil, A. B., Behbehani, A. I.,
and Dashti, H. (2002). Trace elements content in serum, normal skin, and scar
tissues of keloid and normal scar patients. J. Trace Elem. Exp. Med. 15, 57–66.
Bayat, A., Arscott, G., Ollier, W. E. R., McGrouther, D. A., and Ferguson, M. W. J.
(2005). Keloid disease: clinical relevance of single versus multiple site scars. Br.
J. Plast. Surg. 58, 28–37. doi: 10.1016/j.bjps.2004.04.024
Beer, T. W., Baldwin, H., Goddard, J. R., Gallagher, P. J., and Wright, D. H. (1998).
Angiogenesis in pathological and surgical scars. Hum. Pathol. 29, 1273–1278.
Bella, H., Heise, M., Yagi, K. I., Black, G., McGrouther, D. A., and Bayat, A.
(2011). A clinical characterization of familial keloid disease in unique African
tribes reveals distinct keloid phenotypes. Plast. Reconstr. Surg. 127, 689–702.
doi: 10.1097/PRS.0b013e3181fed645
Berman, B., and Duncan, M. R. (1989). Short-term keloid treatment in vivo with
human interferon alfa-2b results in a selective and persistent normalization
of keloidal fibroblast collagen, glycosaminoglycan, and collagenase production
in vitro. J. Am. Acad. Dermatol. 21, 694–702. doi: 10.1016/s0190-9622(89)
70239-5
Berstein, S. C., and Roenigk, R. K. (1996). “Chapter 36, Keloids, in Roenigk
Roenigk’s Dermatologic Surgery Principles and Practice, 2nd Edn, eds R. K.
Roenigk and H. H. Roenigk (New York, NY: Illust. Marcel Dekker, Inc),
603–604.
Bertheim, U., and Hellström, S. (1994). The distribution of hyaluronan in human
skin and mature, hypertrophic and keloid scars. Br. J. Plast. Surg. 47, 483–489.
doi: 10.1016/0007-1226(94)90031- 0
Bettinger, D. A., Yager, D. R., Diegelmann, R. F., and Cohen, I. K. (1996). The
effect of TGF-βon keloid fibroblast proliferation and collagen synthesis. Plast.
Reconstr. Surg. 98, 827–833. doi: 10.1097/00006534- 199610000-00012
Frontiers in Cell and Developmental Biology | www.frontiersin.org 18 May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 19
Limandjaja et al. The Keloid Disorder
Bijlard, E., Kouwenberg, C. A. E., Timman, R., Hovius, S. E. R., Busschbach,
J. J. V., and Mureau, M. A. M. (2017). Burden of keloid disease: a cross-sectional
health-related quality of life assessment. Acta Derm. Venereol. 97, 225–229.
doi: 10.2340/00015555-2498
Blaži´
c, T. M., and Brajac, I. (2006). Defective induction of senescence during
wound healing is a possible mechanism of keloid formation. Med. Hypotheses
66, 649–652. doi: 10.1016/j.mehy.2005.09.033
Bloch, E. F., Hall, M. C., Denson, M. J., and Slay-Solomon, V. (1984).
General immunoreactivity in keloid patients. Plast. Reconstr. Surg. 73,
448–451.
Bloom, D. (1956). Heredity of keloids; review of the literature and report of
a family with multiple keloids in five generations. N. Y. State J. Med. 56,
511–519.
Bloor, B. K., Tidman, N., Leigh, I. M., Odell, E., Dogan, B., Wollina, U., et al.
(2003). Expression of keratin K2e in cutaneous and oral lesions: association
with keratinocyte activation, proliferation, and keratinization. Am. J. Pathol.
Am. Soc. Invest. Pathol. 162, 963–975. doi: 10.1016/S0002-9440(10)63891- 6
Blume-Peytavi, U., Geilen, C. C., Sommer, C., Almond-Roesler, B., and Orfanos,
C. E. (1997). The phospholipid analogue hexadecylphosphocholine (HePC)
inhibits proliferation of keloid fibroblasts in vitro and modulates their
fibronectin and integrin synthesis. Arch. Dermatol. Res. 289, 164–169. doi:
10.1007/s004030050173
Bock, O., Yu, H., Zitron, S., Bayat, A., Ferguson, M. W. J., and Mrowietz, U. (2005).
Studies of transforming growth factors beta 1-3 and their receptors I and II in
fibroblast of keloids and hypertrophic scars. Acta Derm. Venereol. 85, 216–220.
doi: 10.1080/00015550410025453
Boyce, D. E., Ciampolini, J., Ruge, F., Harding, K. G., and Murison, M. M. S. C.
(2001). Inflammatory cell subpopulations in keloid scars. Br. J. Plast. Surg. 54,
511–516. doi: 10.1054/bjps.2001.3638
Bran, G. M., Goessler, U. R., Baftiri, A., Hormann, K., Riedel, F., and Sadick, H.
(2010). Effect of transforming growth factor-β1 antisense oligonucleotides on
matrix metalloproteinases and their inhibitors in keloid fibroblasts. Otolaryngol.
Head Neck Surg. 143, 66–71. doi: 10.1016/j.otohns.2010.03.029
Bran, G. M., Goessler, U. R., Hormann, K., Riedel, F., and Sadick, H. (2009).
Keloids: current concepts of pathogenesis (Review). Int. J. Mol. Med. 24,
283–293. doi: 10.3892/ijmm_00000231
Breasted, J. H. (1930). Case Forty-five, Bulging Tumors on the Breast. Edwin
Smith Surgical Papyrus, Vol. 1, Hieroglphic Transliteration, Translation and
Commentary. Chicago, IL: The University of Chicago Press, 463.
Brody, G. S. (1990). Keloids and hypertrophic scars. Plast. Reconstr. Surg. 86:804.
Broughton, G. II, Janis, J. E., and Attinger, C. E. (2006). The basic science of wound
healing. Plast. Reconstr. Surg 117(Suppl. 7), 12S–34S.
Bucala, R., Spiegel, L. A., Chesney, J., Hogan, M., and Cerami, A. (1994).
Circulating fibrocytes define a new leukocyte subpopulation that mediates tissue
repair. Mol. Med. 1, 71–81.
Burd, A. (2006). Keloid epidemiology: population based studies needed. J. Plast.
Reconstr. Aesthetic Surg. 59, 105–111. doi: 10.1016/j.bjps.2005.07.012
Burd, A., and Chan, E. (2002). Keratinocyte-keloid interaction. Plast. Reconstr.
Surg. 110, 197–202.
Burd, A., and Huang, L. (2005). Hypertrophic response and keloid diathesis: two
very different forms of scar. Plast. Reconstr. Surg. 116, 150e–157e. doi: 10.1097/
01.prs.0000191977.51206.43
Butler, P. D., Longaker, M. T., and Yang, G. P. (2008a). Current progress in
keloid research and treatment. J. Am. Coll. Surg. 206, 731–741. doi: 10.1016/
j.jamcollsurg.2007.12.001
Butler, P. D., Ly, D. P., Longaker, M. T., and Yang, G. P. (2008b). Use of organotypic
coculture to study keloid biology. Am. J. Surg. 195, 144–148. doi: 10.1016/j.
amjsurg.2007.10.003
Butzelaar, L., Niessen, F. B., Talhout, W., Schooneman, D. P. M., Ulrich, M. M.,
Beelen, R. H. J., et al. (2017). Different properties of skin of different body
sites: the root of keloid formation? Wound Repair Regen. 25, 758–766. doi:
10.1111/wrr.12574
Bux, S., and Madaree, A. (2010). Keloids show regional distribution of proliferative
and degenerate connective tissue elements. Cells Tissues Organs 191, 213–234.
doi: 10.1159/000231899
Bux, S., and Madaree, A. (2012). Involvement of upper torso stress amplification,
tissue compression and distortion in the pathogenesis of keloids. Med.
Hypotheses 78, 356–363. doi: 10.1016/j.mehy.2011.12.008
Carrino, D. A., Mesiano, S., Barker, N. M., Hurd, W. W., and Caplan, A. I.
(2012). Proteoglycans of uterine fibroids and keloid scars: similarity in their
proteoglycan composition. Biochem. J. 443, 361–368. doi: 10.1042/BJ20111996
Carroll, L. A., Hanasono, M. M., Mikulec, A. A., Kita, M., and Koch, J. R. (2002).
Triamcinolone stimulates bFGF production and inhibits TGF-β1 production by
human dermal fibroblasts. Dermatol. Surg. 28, 704–709. doi: 10.1046/j.1524-
4725.2002.02012.x
Carroll, L. A., and Koch, R. J. (2003). Heparin stimulates production of bFGF and
TGF-β1 by human normal, keloid, and fetal dermal fibroblasts. Med. Sci. Monit.
9, BR97–BR108.
Chen, Z. D., Zhou, L., Won, T., Gao, Z., Wu, X., and Lu, L. (2018). Characterization
of CD45RO+ memory T lymphocytes in keloid disease.Br. J. Dermatol. 178,
940–950. doi: 10.1111/bjd.16173
Chin, G. S., Liu, W., Peled, Z., Lee, T. Y., Steinbrech, D. S., Hsu, M., et al. (2001).
Differential expression of transforming growth factor-βreceptors I and II and
activation of Smad 3 in keloid fibroblasts. Plast. Reconstr. Surg. 108, 423–429.
doi: 10.1097/00006534-200108000- 00022
Chipev, C. C., Simman, R., Hatch, G., Katz, A. E., Siegel, D. M., and Simon, M.
(2000). Myofibroblast phenotype and apoptosis in keloid and palmar fibroblasts
in vitro. Cell Death Differ. 7, 166–176. doi: 10.1038/sj.cdd.4400605
Chipev, C. C., and Simon, M. (2002). Phenotypic differences between dermal
fibroblasts from different body sites determine their response to tension and
TGFβ1. BMC Dermatol. 2:13. doi: 10.1186/1471-5945-2- 13
Chong, Y., Kim, C. W., Kim, Y. S., Chang, C. H., and Park, T. H. (2018). Complete
excision of proliferating core in auricular keloids significantly reduces local
recurrence: a prospective study. J. Dermatol. 45, 139–144. doi: 10.1111/1346-
8138.14110
Chong, Y., Park, T. H., Seo, S. W., and Chang, C. H. (2015). Histomorphometric
analysis of collagen architecture of auricular keloids in an Asian population.
Dermatol. Surg. 41, 415–422. doi: 10.1097/DSS.0000000000000176
Chua, A. W. C., Ma, D., Gan, S. U., Fu, Z., Han, H. C., Song, C., et al.
(2011). The role of R-spondin2 in keratinocyte proliferation and epidermal
thickening in keloid scarring. J. Invest. Dermatol. 131, 644–654. doi: 10.1038/jid.
2010.371
Clore, J. N., Cohen, I. K., and Diegelmann, R. F. (1979). Quantitative assay of types I
and III collagen synthesized by keloid biopsies and fibroblasts. Biochim. Biophys.
Acta 586, 384–390. doi: 10.1016/0304- 4165(79)90107-7
Cobbold, C. A. (2001). The role of nitric oxide in the formation of keloid and
hypertrophic lesions. Med. Hypotheses 57, 497–502. doi: 10.1054/mehy.2001.
1373
Concannon, M. J., Barrett, B. B., Adelstein, E. H., Thornton, W. H., and Puckett,
C. L. (1993). The inhibition of fibroblast proliferation by a novel monokine:
an in vitro and in vivo study. J. Burn Care Rehabil. 14(2 Pt 1), 141–147. doi:
10.1097/00004630-199303000- 00003
Conway, H., Gillette, R., Smith, J. W., and Findley, A. (1960). Differential diagnosis
of keloids and hypertrophic scars by tissue culture technique with notes on
therapy of keloids by surgical excision and decadron. Plast. Reconstr. Surg. 25,
117–132. doi: 10.1097/00006534- 196002000-00001
Cooke, G. L., Chien, A., Brodsky, A., and Lee, R. C. (2005). Incidence
of hypertrophic scars among African Americans linked to vitamin D-3
metabolism? J. Natl. Med. Assoc. 97, 1004–1009.
Cosman, B., Crikelair, G. F., Ju, D. M., Gaulin, J. C., and Lattes, R. (1961). The
surgical treatment of keloids. Plast. Reconstr. Surg. 27, 335–358.
Daian, T., Ohtsuru, A., Rogounovitch, T., Ishihara, H., Hirano, A., Akiyama-
uchida, Y., et al. (2003). Insulin-like growth factor-I enhances transforming
growth factor-β-induced extracellular matrix protein production through the
P38/activating transcription factor-2 signaling pathway in keloid fibroblasts.
J. Invest. Dermatol. 132, 956–962. doi: 10.1046/j.1523-1747.2003.12143.x
Dalkowski, A., Schuppan, D., Orfanos, C. E., and Zouboulis, C. C. (1999). Increased
expression of tenascin C by keloids in vivo and in vitro. Br. J. Dermatol. 141,
50–56. doi: 10.1046/j.1365- 2133.1999.02920.x
De Felice, B., Wilson, R. R., and Nacca, M. (2009). Telomere shortening may be
associated with human keloids. BMC Med. Genet. 10:110. doi: 10.1186/1471-
2350-10- 110
Delpech, L. (1881). De la Chéloïde et de Son Traitement, 1st Edn. (Paris: A. Parent
(imprimeur de la faculté de médecine)).
Dienus, K., Bayat, A., Gilmore, B. F., and Seifert, O. (2010). Increased expression
of fibroblast activation protein-alpha in keloid fibroblasts: implications for
Frontiers in Cell and Developmental Biology | www.frontiersin.org 19 May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 20
Limandjaja et al. The Keloid Disorder
development of a novel treatment option. Arch. Dermatol. Res. 302, 725–731.
doi: 10.1007/s00403-010- 1084-x
Do, D. V., Ong, C., Khoo, Y., Carbone, A., Lim, C., Wang, S., et al. (2012).
Interleukin-18 system plays an important role in keloid pathogenesis via
epithelial-mesenchymal interactions. Br. J. Dermatol. 166, 1275–1288. doi: 10.
1111/j.1365-2133.2011.10721.x
Dohi, T., Padmanabhan, J., Akaishi, S., Than, P. A., Terashima, M., Matsumoto,
N. N., et al. (2019). The interplay of mechanical stress, strain, and stiffness
at the keloid periphery correlates with increased caveolin-1/ROCK signaling
and scar progression. Plast. Reconstr. Surg. 144, 58–67. doi: 10.1097/PRS.
0000000000005718
Dong, X., Mao, S., and Wen, H. (2013). Upregulation of proinflammatory genes
in skin lesions may be the cause of keloid formation (Review). Biomed. Rep. 1,
833–836. doi: 10.3892/br.2013.169
Drummond, P. D., Dawson, L. F., Wood, F. M., and Fear, M. W. (2017). Up-
regulation of α1-adrenoceptors in burn and keloid scars. Burns 44, 582–588.
doi: 10.1016/j.burns.2017.09.010
Duong, H. S., Zhang, Q., Kobi, A., Le, A., and Messadi, D. V. (2006). Assessment
of morphological and immunohistological alterations in long-term keloid skin
explants. Cells Tissues Organs 181, 89–102. doi: 10.1159/000091098
Dustan, H. P. (1995). Does keloid pathogenesis hold the key to understanding
black/white differences in hypertension severity? Hypertension 26, 858–862.
doi: 10.1161/01.hyp.26.6.858
Ehrlich, H. P., Desmoulière, A., Diegelmann, R. F., Cohen, I. K., Compton, C. C.,
Garner, W. L., et al. (1994). Morphological and immunochemical differences
between keloid and hypertrophic scar. Am. J. Pathol. 145, 105–113.
English, R. S., and Shenefelt, P. D. (1999). Keloids and hypertrophic scars.
Dermatol. Surg. 25, 631–638.
Erdag, G., Qureshi, H. S., Patterson, J. W., and Wick, M. R. (2008). CD34-positive
dendritic cells disappear from scars but are increased in pericicatricial tissue.
J. Cutan. Pathol. 35, 752–756. doi: 10.1111/j.1600-0560.2007.00895.x
Fang, F., Huang, R., Zheng, Y., Liu, M., and Huo, R. (2016). Bone marrow derived
mesenchymal stem cells inhibit the proliferative and pro brotic phenotype of
hypertrophic scar fibroblasts and keloid fibroblasts through paracrine signaling.
J. Dermatol. Sci. 83, 95–105. doi: 10.1016/j.jdermsci.2016.03.003
Fong, E. P., and Bay, B. H. (2002). Keloids - the sebum hypothesis revisited. Med.
Hypotheses 58, 264–269. doi: 10.1054/mehy.2001.1426
Fong, E. P., Chye, L. T., and Tan, W. T. L. (1999). Keloids: time to dispel the myths?
Plast. Reconstr. Surg. 104, 1199–1201.
Friedman, D. W., Boyd, C. D., Mackenzie, J. W., Norton, P., Olson, R. M., and
Deak, S. B. (1993). Regulation of collagen gene expression in keloids and
hypertrophic scars. J. Surg. Res. 55, 214–222. doi: 10.1006/jsre.1993.1132
Fujiwara, M., Muragaki, Y., and Ooshima, A. (2005a). Keloid-derived fibroblasts
show increased secretion of factors involved in collagen turnover and depend
on matrix metalloproteinase for migration. Br. J. Dermatol. 153, 295–300. doi:
10.1111/j.1365-2133.2005.06698.x
Fujiwara, M., Muragaki, Y., and Ooshima, A. (2005b). Upregulation of
transforming growth factor-β1 and vascular endothelial growth factor in
cultured keloid fibroblasts: relevance to angiogenic activity. Arch. Dermatol. Res.
297, 161–169. doi: 10.1007/s00403- 005-0596-2
Funayama, E., Chodon, T., Oyama, A., and Sugihara, T. (2003). Keratinocytes
promote proliferation and inhibit apoptosis of the underlying fibroblasts: an
important role in the pathogenesis of keloid. J. Invest. Dermatol. 121, 1326–
1331. doi: 10.1111/j.1523- 1747.2003.12572.x
Furtado, F., Hochman, B., Farber, P. L., Muller, M. C., Hayashi, L. F.,
and Ferreira, L. M. (2012). Psychological stress as a risk factor for
postoperative keloid recurrence. J. Psychosom. Res. 72, 282–287. doi:
10.1016/j.jpsychores.2011.12.010
Gao, F. L., Jin, R., Zhang, L., and Zhang, Y. G. (2013). The contribution of
melanocytes to pathological scar formation during wound healing. Int. J. Clin.
Exp. Med. 6, 609–613.
Ghazizadeh, M., Tosa, M., Shimizu, H., Hyakusoku, H., and Kawanami, O. (2007).
Functional implications of the IL-6 signaling pathway in keloid pathogenesis.
J. Invest. Dermatol. 127, 98–105. doi: 10.1038/sj.jid.5700564
Giugliano, G., Pasquali, D., Notaro, A., Brongo, S., Nicoletti, G., D’Andrea, F.,
et al. (2003). Verapamil inhibits interleukin-6 and vascular endothelial growth
factor production in primary cultures of keloid fibroblasts. Br. J. Plast. Surg. 56,
804–809. doi: 10.1016/s0007- 1226(03)00384-9
Glass, I. I. D. A. (2017). Current understanding of the genetic causes of keloid
formation. J. Investig. Dermatol. Symp. Proc. 18, S50–S53. doi: 10.1016/j.jisp.
2016.10.024
Gold, M. H., Berman, B., Clementoni, M. T., Gauglitz, G. G., Nahai, F., and
Murcia, C. (2014). Updated international clinical recommendations on scar
management: part 1 - evaluating the evidence. Dermatol. Surg. 40, 817–824.
doi: 10.1111/dsu.0000000000000049
Granick, M., Kimura, M., Kim, S., Daniali, L., Cao, X., Herbig, U., et al. (2011).
Telomere dynamics in keloids. Eplasty 11:e15.
Guadanhim, L. R. S., Gonçalves, R. G., and Bagatin, E. (2016). Observational
retrospective study evaluating the effects of oral isotretinoin in keloids and
hypertrophic scars. Int. J. Dermatol. 55, 1255–1258. doi: 10.1111/ijd.13317
Gulamhuseinwala, N., Mackey, S., Meagher, P., and Powell, B. (2008). Should
excised keloid scars be sent for routine histologic analysis? Ann. Plast. Surg. 60,
186–187. doi: 10.1097/SAP.0b013e318056d6cc
Hahn, J. M., Glaser, K., McFarland, K. L., Aronow, B. J., Boyce, S. T., and
Supp, D. M. (2013). Keloid-derived keratinocytes exhibit an abnormal gene
expression profile consistent with a distinct causal role in keloid pathology.
Wound Repair Regen. 21, 530–544. doi: 10.1111/wrr.12060
Hahn, J. M., Mcfarland, K. L., Combs, K. A., and Supp, D. M. (2016).
Partial epithelial-mesenchymal transition in keloid scars: regulation of keloid
keratinocyte gene expression by transforming growth factor-β1. Burns Trauma
4:30. doi: 10.1186/s41038- 016-0055-7
Haisa, M., Okochi, H., and Grotendorst, G. R. (1994). Elevated levels of PDGF-α
receptors in keloid fibroblasts contribute to an enhanced response to PDGF.
J. Invest. Dermatol. 103, 560–563. doi: 10.1111/1523- 1747.ep12396856
Hanasono, M. M., Kita, M., Mikulec, A. A., Lonergan, D., and Koch, R. J.
(2003). Autocrine growth factor production by fetal, keloid, and normal dermal
fibroblasts. Arch. Facial Plast. Surg. 5, 26–30. doi: 10.1001/archfaci.5.1.26
Hanasono, M. M., Lum, J., Carroll, L. A., Mikulec, A. A., and Koch, R. J. (2004).
The effect of silicone gel on basic fibroblast growth factor levels in fibroblast cell
culture. Arch. Facial Plast. Surg. 6, 88–93. doi: 10.1001/archfaci.6.2.88
Hasegawa, T., Nakao, A., Sumiyoshi, K., Tsuboi, R., and Ogawa, H. (2003). IFN-γ
fails to antagonize fibrotic effect of TGF-βon keloid-derived dermal fibroblasts.
J. Dermatol. Sci. 32, 19–24. doi: 10.1016/s0923- 1811(03)00044-6
He, S., Yang, Y., Liu, X., Huang, W., Zhang, X., Yang, S., et al. (2012). Compound
Astragalus and Salvia miltiorrhiza extract inhibits cell proliferation, invasion
and collagen synthesis in keloid fibroblasts by mediating transforming growth
factor-β/Smad pathway. Br. J. Dermatol. 166, 564–574. doi: 10.1111/j.1365-
2133.2011.10674.x
He, Y., Deng, Z., Alghamdi, M., Lu, L., Fear, M. W., and He, L. (2017). From
genetics to epigenetics: new insights into keloid scarring. Cell Prolif. 50:e12326.
doi: 10.1111/cpr.12326
Hellström, M., Hellström, S., Engström-Laurent, A., and Bertheim, U. (2014). The
structure of the basement membrane zone differs between keloids, hypertrophic
scars and normal skin: a possible background to an impaired function. J. Plast.
Reconstr. Aesthetic Surg. 67, 1564–1572. doi: 10.1016/j.bjps.2014.06.014
Hillmer, M. P., and Macleod, S. M. (2002). Experimental keloid scar models: a
review of methodological issues. J. Cutan. Med. Surg. 6, 354–359. doi: 10.1007/
s10227-001- 0121-y
Hochman, B., Isoldi, F. C., Furtado, F., and Ferreira, L. M. (2015). New approach
to the understanding of keloid: psychoneuroimmune–endocrine aspects. Clin.
Cosmet. Investig. Dermatol. 8, 67–73. doi: 10.2147/CCID.S49195
Hochman, B., Nahas, F. X., Sobral, C. S., Arias, V., Locali, R. F., Juliano, Y., et al.
(2008). Nerve fibres: a possible role in keloid pathogenesis. Br. J. Dermatol. 158,
624–657. doi: 10.1111/j.1365- 2133.2007.08401.x
Hollywood, K. A., Maatje, M., Shadi, I. T., Henderson, A., McGrouther, D. A.,
Goodacre, R., et al. (2010). Phenotypic profiling of keloid scars using FT-IR
microspectroscopy reveals a unique spectral signature. Arch. Dermatol. Res. 302,
705–715. doi: 10.1007/s00403-010- 1071-2
Hsu, C., Lin, H., Harn, H. I., Ogawa, R., Wang, Y., Ho, Y., et al. (2018).
Caveolin-1 controls hyperresponsiveness to mechanical stimuli and activation
in keloid fibroblasts. J. Invest. Dermatol. 138, 208–218. doi: 10.1016/j.jid.2017.
05.041
Hsu, Y. C., Hsiao, M., Wang, L. F., Chien, Y. W., and Lee, W. R. (2006). Nitric
oxide produced by iNOS is associated with collagen synthesis in keloid scar
formation. Nitric Oxide Biol. Chem. 14, 327–334. doi: 10.1016/j.niox.2006.
01.006
Frontiers in Cell and Developmental Biology | www.frontiersin.org 20 May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 21
Limandjaja et al. The Keloid Disorder
Hu, Z., Tang, B., Guo, D., Zhang, J., Liang, Y., Ma, D., et al. (2014). Expression of
insulin-like growth factor-1 receptor in keloid and hypertrophic scar. Clin. Exp.
Dermatol. 39, 822–828. doi: 10.1111/ced.12407
Huang, C., Akaishi, S., Hyakusoku, H., and Ogawa, R. (2014). Are keloid and
hypertrophic scar different forms of the same disorder? A fibroproliferative
skin disorder hypothesis based on keloid findings. Int. Wound J. 11, 517–522.
doi: 10.1111/j.1742-481X.2012.01118.x
Huang, C., Liu, L., You, Z., Wang, B., Du, Y., and Ogawa, R. (2017). Keloid
progression: a stiffness gap hypothesis. Int. Wound J. 14, 764–771. doi: 10.1111/
iwj.12693
Huang, C., and Ogawa, R. (2013a). Pharmacological treatment for keloids. Expert
Opin. Pharmacother. 14, 2087–2100.
Huang, C., and Ogawa, R. (2013b). Roles of lipid metabolism in keloid
development. Lipids Health Dis. 12, 1–6. doi: 10.1186/1476-511X-12-60
Huang, C., and Ogawa, R. (2014). The link between hypertension and pathological
scarring: does hypertension cause or promote keloid and hypertrophic scar
pathogenesis? Wound Repair Regen. 22, 462–466. doi: 10.1111/wrr.12197
Hunzelmann, N., Anders, S., Sollberg, S., Schönherr, E., and Krieg, T. (1996). Co-
ordinate induction of collagen type I and biglycan expression in keloids. Br. J.
Dermatol. 135, 394–399.
ICD-10 (2016). World Heal. Organ. [cited 2019 Sep 3]. Available online at: https:
//icd.who.int/browse10/2016/en#/L91.0 (accessed Sep 3, 2019).
ICD-11 (2018). World Heal. Organ. [cited 2019 Sep 3]. Available online at: https:
//www.who.int/classifications/icd/en/ (accessed Sep 3, 2019).
Ikeda, M., Naitoh, M., Kubota, H., Ishiko, T., Yoshikawa, K., Yamawaki, S.,
et al. (2009). Elastic fiber assembly is disrupted by excessive accumulation
of chondroitin sulfate in the human dermal fibrotic disease, keloid.
Biochem. Biophys. Res. Commun. 390, 1221–1228. doi: 10.1016/j.bbrc.2009.
10.125
Iqbal, S. A., Sidgwick, G. P., and Bayat, A. (2012). Identification of fibrocytes
from mesenchymal stem cells in keloid tissue: a potential source of abnormal
fibroblasts in keloid scarring. Arch. Dermatol. Res. 304, 665–671. doi: 10.1007/
s00403-012- 1225-5
Iqbal, S. A., Syed, F., McGrouther, D. A., Paus, R., and Bayat, A. (2010). Differential
distribution of haematopoietic and nonhaematopoietic progenitor cells in
intralesional and extralesional keloid: do keloid scars provide a niche for
nonhaematopoietic mesenchymal stem cells? Br. J. Dermatol. 162, 1377–1383.
doi: 10.1111/j.1365-2133.2010.09738.x
Jfri, A., Rajeh, N., and Karkashan, E. (2015). A case of multiple spontaneous keloid
scars. Case Rep. Dermatol. 7, 156–160. doi: 10.1159/000437249
Jiao, H., Fan, J., Cai, J., Pan, B., Yan, L., Dong, P., et al. (2015). Analysis of
characteristics similar to autoimmune disease in keloid patients. Aesthetic Plast.
Surg. 39, 818–825. doi: 10.1007/s00266-015- 0542-4
Jiao, H., Zhang, T., Fan, J., and Xiao, R. (2017). The superficial dermis may initiate
keloid formation: histological analysis of the keloid dermis at different depths.
Front. Physiol. 8:885. doi: 10.3389/fphys.2017.00885
Jin, Q., Gui, L., Niu, F., Yu, B., Lauda, N., Liu, J., et al. (2018). Macrophages in keloid
are potent at promoting the differentiation and function of regulatory T-cells.
Exp. Cell Res. 362, 472–476. doi: 10.1016/j.yexcr.2017.12.011
Jin, Z. (2014). Increased c-Met phosphorylation is related to keloid pathogenesis:
implications for the biological behaviour of keloid fibroblasts. Pathology 46,
25–31. doi: 10.1097/PAT.0000000000000028
Jing, C., Jia-han, W., and Hong-Xing, Z. (2010). Double-edged effects of
neuropeptide substance P on repair of cutaneous trauma. Wound Repair Regen.
18, 319–324. doi: 10.1111/j.1524-475X.2010.00589.x
Jumper, N., Hodgkinson, T., Paus, R., and Bayat, A. (2017). A role for Neuregulin-
1 in promoting keloid fibroblast migration. Acta Derm. Venereol. 97, 675–684.
doi: 10.2340/00015555-2587
Jumper, N., Paus, R., and Bayat, A. (2015). Functional histopathology of keloid
disease. Histol. Histopathol. 30, 1033–1057. doi: 10.14670/HH-11-624
Jurzak, M., and Adamczyk, K. (2013). Influence of genistein on c-Jun, c-Fos and
Fos-B of AP-1 subunits expression in skin keratinocytes, fibroblasts and keloid
fibroblasts cultured in vitro. Acta Pol. Pharm. Drug Res. 70, 205–213.
Jurzak, M., Adamczyk, K., Anto´
nczak, P., Garncarczyk, A., Kus¨
smierz, D.,
and Latocha, M. (2014). Evaluation of genistein ability to modulate CTGF
mRNA/protein expression, genes expression of TGF-βisoforms and expression
of selected genes regulating cell cycle in keloid fibroblasts in vitro. Acta Pol.
Pharm. Drug Res. 71, 972–986.
Kamath, N. V., Ormsby, A., Bergfeld, W. F., and House, N. S. (2002). A light
microscopic. J. Cutan. Pathol. 29, 27–32.
Khoo, Y. T., Ong, C. T., Mukhopadhyay, A., Han, H. C., Do, D. V., Lim, I. J.,
et al. (2006). Upregulation of secretory connective tissue growth factor (CTGF)
in keratinocyte-fibroblast coculture contributes to keloid pathogenesis. J. Cell.
Physiol. 208, 336–343. doi: 10.1002/jcp.20668
Khorshid, F. A. (2005). Comparative study of keloid formation in humans and
laboratory animals. Med. Sci. Monit. 11, BR212–BR219.
Kiprono, S. K., Chaula, B. M., Masenga, J. E., Muchunu, J. W., Mavura, D. R., and
Moehrle, M. (2015). Epidemiology of keloids in normally pigmented Africans
and African people with albinism: population-based cross-sectional survey. Br.
J. Dermatol. 173, 852–854. doi: 10.1111/bjd.13826
Kischer, C. (1984). Comparative ultrastructure of hypertrophic scars and keloids.
Scan. Electron Microsc. (Pt 1), 423–431.
Kischer, C. W., and Hendrix, M. J. C. (1983). Fibronectin (FN) in hypertrophic
scars and keloids. Cell Tissue Res. 231, 29–37. doi: 10.1007/bf00215771
Kischer, C. W., and Pindur, J. (1990). Effects of platelet derived growth factor
(PDGF) on fibronectin (FN) production by human skin and scar fibroblasts.
Cytotechnology 3, 231–238. doi: 10.1007/BF00365486
Kischer, C. W., Thies, A. C., and Chvapil, M. (1982). Perivascular myofibroblasts
and microvascular occlusion in hypertrophic scars and keloids. Hum. Pathol.
13, 819–824. doi: 10.1016/s0046- 8177(82)80078-6
Koonin, A. J. (1964). The aetiology of keloids: a review of the literature and a new
hypothesis. S. Afr. Med. J. 38, 913–916.
Kroeze, K. L., Jurgens, W. J., Doulabi, B. Z., van Milligen, F. J., Scheper, R. J.,
and Gibbs, S. (2009). Chemokine-mediated migration of skin-derived stem
cells: predominant role for CCL5/RANTES. J. Invest. Dermatol. 129, 1569–1581.
doi: 10.1038/jid.2008.405
Kurokawa, N., Ueda, K., and Tsuji, M. (2010). Study of microvascular structure
in keloid and hypertrophic scars: density of microvessels and the efficacy of
three-dimensional vascular imaging. J. Plast. Surg. Hand Surg. 44, 272–277.
doi: 10.3109/2000656X.2010.532923
Kuwahara, H., Tosa, M., Murakami, M., Mohammad, G., and Ogawa, R.
(2016). Examination of epithelial mesenchymal transition in keloid tissues and
possibility of keloid therapy target. Plast. Reconstr. Surg. Glob. Open 4:e1138.
doi: 10.1097/GOX.0000000000001138
Ladin, D. A., Hou, Z., Patel, D., McPhail, M., Olson, J. C., Saed, G. M., et al. (1998).
P53 and apoptosis alterations in keloids and keloid fibroblasts. Wound Repair
Regen. 6, 28–37. doi: 10.1046/j.1524- 475x.1998.60106.x
Le, A. D., Zhang, Q., Wu, Y., Messadi, D. V., Akhondzadeh, A., Nguyen, A. L.,
et al. (2004). Elevated vascular endothelial growth factor in keloids: relevance to
tissue fibrosis. Cells Tissues Organs 176, 87–94. doi: 10.1159/000075030
Lebeko, M., Khumalo, N. P., and Bayat, A. (2019). Multi-dimensional models
for functional testing of keloid scars: in silico, in vitro, organoid, organotypic,
ex vivo organ culture, and in vivo models. Wound Repair Regen. 27, 298–308.
doi: 10.1111/wrr.12705
Lee, C. H., Hong, C. H., Chen, Y. T., Chen, Y. C., and Shen, M. R. (2012). TGF-
beta1 increases cell rigidity by enhancing expression of smooth muscle actin:
keloid-derived fibroblasts as a model for cellular mechanics. J. Dermatol. Sci.
67, 173–180. doi: 10.1016/j.jdermsci.2012.06.004
Lee, J. H., Shin, J. U., Jung, I., Lee, H., Rah, D. K., Jung, J. Y., et al. (2013). Proteomic
profiling reveals upregulated protein expression of Hsp70 in keloids. Biomed
Res. Int. 2013:621538. doi: 10.1155/2013/621538
Lee, J. Y. Y., Yang, C. C., Chao, S. C., and Wong, T. W. (2004). Histopathological
differential diagnosis of keloid and hypertrophic scar. Am. J. Dermatopathol. 26,
379–384. doi: 10.1097/00000372- 200410000-00006
Lee, K. S., Song, J. Y., and Suh, M. H. (1991). Collagen mRNA expression detected
by in situ hybridization in keloid tissue. J. Dermatol. Sci. 2, 316–323. doi:
10.1016/0923-1811(91)90056- 4
Lee, S. S., Yosipovitch, G., Chan, Y. H., and Goh, C. L. (2004). Pruritus, pain, and
small nerve fiber function in keloids: a controlled study. J. Am. Acad. Dermatol.
51, 1002–1006. doi: 10.1016/j.jaad.2004.07.054
Lee, W. J., Park, J. H., Shin, J. U., Noh, H., Lew, D. H., Yang, W. I.,
et al. (2015). Endothelial-to-mesenchymal transition induced by Wnt3a in
keloid pathogenesis. Wound Repair Regen. 23, 435–442. doi: 10.1111/wrr.
12300
Lee, Y.-S., Hsu, T., Chiu, W.-C., Sarkozy, H., Kulber, D. A., Choi, A., et al. (2015).
Keloid-derived, plasma/fibrin-based skin equivalents generate de novo dermal
Frontiers in Cell and Developmental Biology | www.frontiersin.org 21 May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 22
Limandjaja et al. The Keloid Disorder
and epidermal pathology of keloid fibrosis in a mouse model. Wound Repair
Regen. 24, 302–316. doi: 10.1111/wrr.12397
Li, M., and Wu, L. (2016). Functional analysis of keratinocyte and fibroblast
gene expression in skin and keloid scar tissue based on deviation analysis of
dynamic capabilities. Exp. Ther. Med. 12, 3633–3641. doi: 10.3892/etm.2016.
3817
Liang, C. J., Yen, Y. H., Hung, L. Y., Wang, S. H., Pu, C. M., Chien,
H. F., et al. (2013). Thalidomide inhibits fibronectin production in TGF-
β1-treated normal and keloid fibroblasts via inhibition of the p38/SMAD3
pathway. Biochem. Pharmacol. 85, 1594–1602. doi: 10.1016/j.bcp.2013.
02.038
Lim, C. K., Halim, A. S., Yaacob, N. S., Zainol, I., and Noorsal, K. (2013).
Keloid pathogenesis via Drosophila similar to mothers against decapentaplegic
(SMAD) signaling in a primary epithelial-mesenchymal in vitro model treated
with biomedical-grade chitosan porous skin regenerating template. J. Biosci.
Bioeng. 115, 453–458. doi: 10.1016/j.jbiosc.2012.10.010
Lim, C. P., Phan, T. T., Lim, I. J., and Cao, X. (2006). Stat3 contributes to
keloid pathogenesis via promoting collagen production, cell proliferation and
migration. Oncogene 25, 5416–5425. doi: 10.1038/sj.onc.1209531
Lim, C. P., Phan, T. T., Lim, I. J., and Cao, X. (2009). Cytokine profiling and
Stat3 phosphorylation in epithelial-mesenchymal interactions between keloid
keratinocytes and fibroblasts. J. Invest. Dermatol. 129, 851–861. doi: 10.1038/
jid.2008.337
Lim, I. J., Phan, T. T., Bay, B. H., Qi, R., Huynh, H., Tan, W. T. L., et al. (2002).
Fibroblasts cocultured with keloid keratinocytes: normal fibroblasts secrete
collagen in a keloidlike manner. Am. J. Physiol. Cell Physiol. 283, C212–C222.
doi: 10.1152/ajpcell.00555.2001
Lim, I. J., Phan, T. T., Song, C., Tan, W. T. L., and Longaker, M. T. (2001).
Investigation of the influence of keloid-derived keratinocytes on fibroblast
growth and proliferation in vitro. Plast. Reconstr. Surg. 107, 787–808. doi:
10.1097/00006534-200103000- 00022
Lim, I. J., Phan, T. T., Tan, E. K., Nguyen, T. T. T., Tran, E., Longaker, M. T.,
et al. (2003). Synchronous activation of ERK and phosphatidylinositol 3-kinase
pathways is required for collagen and extracellular matrix production in keloids.
J. Biol. Chem. 278, 40851–40858. doi: 10.1074/jbc.M305759200
Lim, K. H., Itinteang, T., Davis, P. F., and Tan, S.T. (2019). Experimental stem cells
in keloid lesions: a review. Plast. Reconstr. Surg. Glob. Open 5, 1–6.
Limandjaja, G. C. (2019). The Keloid Disorder: Histopathology and In Vitro
Reconstruction, Doctoral Dissertation, 1st Edn. Rotterdam: Optima Grafische
Communicatie.
Limandjaja, G. C., Belien, J. M., Scheper, R. J., Niessen, F. B., and Gibbs, S.
(2019). Hypertrophic and keloid scars fail to progress from the CD34-/α-
smooth muscle actin (α-SMA)+ immature scar phenotype and show gradient
differences in α-SMA and p16 expression. Br. J. Dermatol. 182, 974–986. doi:
10.1111/bjd.18219
Limandjaja, G. C., van den Broek, L. J., Breetveld, M., Waaijman, T., Monstrey,
S., De Boer, E. M., et al. (2018a). Characterization of in vitro reconstructed
human normotrophic, hypertrophic, and keloid scar models. Tissue Eng. Part
C Methods 24, 242–253. doi: 10.1089/ten.TEC.2017.0464
Limandjaja, G. C., van den Broek, L. J., Waaijman, T., van Veen, H. A., Everts,
V., Monstrey, S., et al. (2017). Increased epidermal thickness and abnormal
epidermal differentiation in keloid scars. Br. J. Dermatol. 176, 116–126. doi:
10.1111/bjd.14844
Limandjaja, G. C., van den Broek , L. J., Waaijman, T., Breetveld, M., and Monstrey,
S. (2018b). Reconstructed human keloid models show heterogeneity within
keloid scars. Arch. Dermatol. Res. 310, 815–826. doi: 10.1007/s00403-018-
1873-1
Lin, L., Wang, Y., Liu, W., and Huang, Y. (2015). BAMBI inhibits skin fibrosis
in keloid through suppressing TGF-β1-induced hypernomic fibroblast cell
proliferation and excessive accumulation of collagen I. Int. J. Clin. Exp. Med.
8, 13227–13234.
Liu, Q., Wang, X., Jia, Y., Long, X., Yu, N., Wang, Y., et al. (2016). Increased blood
flow in keloids and adjacent skin revealed by laser speckle contrast imaging.
Lasers Surg. Med. 48, 360–364. doi: 10.1002/lsm.22470
Lorenz, H. P., and Longaker, M. T. (2003). “Wounds: biology, pathology, and
management, in Essent. Pract. Surg, eds M. Li, J. Norton, R. R. Bollinger, A. E.
Chang, S. F. Lowry, S. J. Mulvihill, et al. (New York, NY: Springer-Verlag),
77–88.
Lotti, T., Teofoli, P., Bianchi, B., and Mauviel, A. (1999). POMC and fibroblast
biology. Ann. N. Y. Acad. Sci. 885, 262–267. doi: 10.1111/j.1749-6632.1999.
tb08683.x
Louw, L., and Dannhauser, A. (2000). Keloids in rural black South Africans. Part
2: dietary fatty acid intake and total phospholipid fatty acid profile in the
blood of keloid patients. Prostaglandins Leukot. Essent. Fatty Acids 63, 247–253.
doi: 10.1054/plef.2000.0208
Louw, L., van der Westhuizen, J., de Wit, L. D., and Edwards, G. (1997).
Keloids: peripheral and central differences in cell morphology and fatty acid
compositions of lipids. Adv. Exp. Med. Biol. 407, 515–520. doi: 10.1007/978- 1-
4899-1813- 0_77
Lu, F., Gao, J., Ogawa, R., Hyakusoku, H., and Ou, C. (2007a). Biological differences
between fibroblasts derived from peripheral and central areas of keloid tissues.
Plast. Reconstr. Surg. 120, 625–630. doi: 10.1097/01.prs.0000270293.93612.7b
Lu, F., Gao, J., Ogawa, R., Hyakusoku, H., and Ou, C. (2007b). Fas-mediated
apoptotic signal transduction in keloid and hypertrophic scar. Plast. Reconstr.
Surg. 119, 1714–1721. doi: 10.1097/01.prs.0000258851.47193.06
Lu, W. S., Zheng, X. D., Yao, X. H., and Zhang, L. F. (2015). Clinical and
epidemiological analysis of keloids in Chinese patients. Arch. Dermatol. Res.
307, 109–114. doi: 10.1007/s00403-014- 1507-1
Luo, L., Li, J., Liu, H., Jian, X., Zou, Q., Zhao, Q., et al. (2017). Adiponectin is
involved in connective tissue growth factor-induced proliferation, migration
and overproduction of the extracellular matrix in keloid fibroblasts. Int. J. Mol.
Sci. 18, 1–21. doi: 10.3390/ijms18051044
Luo, L. F, Shi, Y., Zhou, Q., Xu, S. Z., and Lei, T. C. (2013). Insufficient expression of
the melanocortin-1 receptor by human dermal fibroblasts contributes to excess
collagen synthesis in keloid scars. Exp. Dermatol. 22, 764–766. doi: 10.1111/exd.
12250
Luo, S., Benathan, M., Raffoul, W., Panizzon, R. G., and Egloff, D. V. (2001).
Abnormal balance between proliferation and apoptotic cell death in fibroblasts
derived from keloid lesions. Plast. Reconstr. Surg. 107, 87–96. doi: 10.1097/
00006534-200101000- 00014
Ma, X., Chen, J., Xu, B., Long, X., Qin, H., Zhao, R. C., et al. (2015). Keloid-
derived keratinocytes acquire a fibroblast-like appearance and an enhanced
invasive capacity in a hypoxic microenvironment in vitro. Int. J. Mol. Med. 35,
1246–1256. doi: 10.3892/ijmm.2015.2135
Maeda, D., Kubo, T., Kiya, K., Kawai, K., Kobayashi, D., Fujiwara, T., et al.
(2019). Periostin is induced by IL-4/IL-13 in dermal fibroblasts and promotes
RhoA/ROCK pathway-mediated TGF-β1 secretion in abnormal scar formation.
J. Plast. Surg. Hand Surg. 53, 288–294. doi: 10.1080/2000656X.2019.1612752
Mancini, R. E., and Quaife, J. V. (1962). Histogenesis of experimentally produced
keloids. J. Invest. Dermatol. 38, 143–181.
Marneros, A. G., and Krieg, T. (2004). Keloids-clinical diagnosis, pathogenesis, and
treatment options. J. Dtsch. Dermatol. Ges. 2, 905–913. doi: 10.1046/j.1439-
0353.2004.04077.x
Marneros, A. G., Norris, J. E. C., Olsen, B. R., and Reichenberger, E. (2001). Clinical
genetics of familial keloids. Arch. Dermatol. 137, 1429–1434.
Marttala, J., Andrews, J. P., Rosenbloom, J., and Uitto, J. (2016). Keloids: animal
models and pathologic equivalents to study tissue fibrosis. Matrix Biol. Int. Soc.
Matrix Biol. 51, 47–54. doi: 10.1016/j.matbio.2016.01.014
Materazzi, S., Pellerito, S., Di Serio, C., Paglierani, M., Naldini, A., Ardinghi, C.,
et al. (2007). Analysis of protease-activated receptor-1 and -2 in human scar
formation. J. Pathol. Engl. 212, 440–449. doi: 10.1002/path.2197
Mathangi Ramakrishnan, K., Meenakshi Janakiraman, M., and Babu, M. (2012).
Expression of fibrocyte markers by keloid fibroblasts: an insight into fibrosis
during burn wound healing a preliminary study. Ann. Burns Fire Disasters 25,
148–151.
McCauley, R. L., Vimlarani, C., Ying-Yue, L., Herndon, D. N., and Robson, M. C.
(1992). Altered cytokines production in black patients with keloids. J. Clin.
Immunol. 12, 300–308. doi: 10.1007/bf00918154
Mccormack, M. C., Nowak, K. C., and Koch, R. J. (2001). The effect of copper
tripeptide and tretinoin on growth factor production in a serum-free fibroblast
model. Arch. Facial Plast. Surg. 3, 28–32.
McCoy, B. J., Galdun, J., and Cohen, K. (1982). Effects of density and cellular aging
on collagen synthesis and growth kinetics in keloid and normal skin fibroblasts.
In Vitro 18, 79–86. doi: 10.1007/bf02796388
McFarland, K. L., Glaser, K., Hahn, J. M., Boyce, S. T., and Supp, D. M. (2011).
Culture medium and cell density impact gene expression in normal skin and
Frontiers in Cell and Developmental Biology | www.frontiersin.org 22 May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 23
Limandjaja et al. The Keloid Disorder
abnormal scar-derived fibroblasts. J. Burn Care Res. 32, 498–508. doi: 10.1097/
BCR.0b013e3182223cb1
Meenakshi, J., Jayaraman, V., Ramakrishnan, K. M., and Babu, M. (2005).
Ultrastructural differentiation of abnormal scars. Ann Burn. Fire Disasters 18,
83–88.
Meenakshi, J., Vidyameenakshi, S., Ananthram, D., Ramakrishnan, K. M.,
Jayaraman, V., and Babu, M. (2009). Low decorin expression along with
inherent activation of ERK1,2 in earlobe keloids. Burns 35, 519–526. doi:
10.1016/j.burns.2008.07.012
Mendoza-Garcia, J., Sebastian, A., Alonso-Rasgado, T., and Bayat, A. (2015).
Ex vivo evaluation of the effect of photodynamic therapy on skin scars and
striae distensae. Photodermatol. Photoimmunol. Photomed. 31, 239–251. doi:
10.1111/phpp.12180
Messadi, D., Doung, H. S., Zhang, Q., Kelly, A. P., Tuan, T. L., Reichenberger, E.,
et al. (2004). Activation of NFκB signal pathways in keloid fibroblasts. Arch.
Dermatol. Res. 296, 125–133.
Messadi, D., Le, A., Berg, S., Jewett, A., Wen, Z., Kelly, P., et al. (1999). Expression
of apoptosis-associated genes by human dermal scar fibroblasts. Wound Repair
Regen. 7, 511–517. doi: 10.1046/j.1524- 475x.1999.00511.x
Messadi, D. V., Le, A., Berg, S., Huang, G., Zhuang, W., and Bertolami, C. N. (1998).
Effect of TGF-β1 on PDGF receptor expression in human scar fibroblasts. Front.
Biosci. 3, a16–a22.
Meyer, L. J. M., Russell, S. B., Russell, J. D., Trupin, J. S., Egbert, B. M., Shuster,
S., et al. (2000). Reduced hyaluronan in keloid tissue and cultured keloid
fibroblasts. J. Invest. Dermatol. 114, 953–959. doi: 10.1046/j.1523- 1747.2000.
00950.x
Middelkoop, E., Monstrey, S., Téot, L., and Vranckx, J. J. (2011). Scar Management:
Practical Guidelines. Elsene: Maca-Cloetens, 11–109.
Mikulec, A. A., Hanasono, M. M., Lum, J., Kadleck, J. M., Kita, M., and Koch, R. J.
(2001). Effect of tamoxifen on transforming growth factor β1 production by
keloid and fetal fibroblasts. Arch. Facial Plast. Surg. 3, 111–114. doi: 10.1001/
archfaci.3.2.111
Miller, C. C., Godeau, G., Lebreton-DeCoster, C., Desmoulière, A., Pellat, B.,
Dubertret, L., et al. (2003). Validation of a morphometric method for evaluating
fibroblast numbers in normal and pathologic tissues. Exp. Dermatol. 12, 403–
411. doi: 10.1034/j.1600- 0625.2003.00023.x
Moon, J., Kwak, S. S., Park, G., Jung, H., Yoon, B. S., Park, J., et al. (2008). Isolation
and characterization of multipotent human keloid-derived mesenchymal-
like stem cells. Stem Cells Dev. 17, 713–724. doi: 10.1089/scd.2007.
0210
Morris, D. E., Wu, L., Zhao, L. L., Bolton, L., Roth, S. I., Ladin, D. A., et al. (1997).
Acute and chronic animal models for excessive dermal scarring: quantitative
studies. Plast. Reconstr. Surg. 100, 674–681. doi: 10.1097/00006534-199709000-
00021
Moshref, S., and Mufti, S. T. (2009). Keloid and hypertrophic scars: comparative
histopathological and immunohistochemical study. J. King Abdulaziz Univ.
Med. Sci. 17, 3–22.
Muir, I. F. K. (1990). On the nature of keloid and hypertrophic scars. Br. J. Plast.
Surg. 43, 61–69.
Mukhopadhyay, A., Fan, S., Dang, V. D., Khoo, A., Ong, C. T., Lim, I. J.,
et al. (2010). The role of hepatocyte growth factor/c-Met system in keloid
pathogenesis. J. Trauma Inj. Infect. Crit. Care 69, 1457–1466. doi: 10.1097/TA.
0b013e3181f45f71
Murao, N., Seino, K.-I., Hayashi, T., Ikeda, M., Funayama, E., Furukawa,
H., et al. (2014). Treg-enriched CD4+ T cells attenuate collagen synthesis
in keloid fibroblasts. Exp. Dermatol. 23, 266–271. doi: 10.1111/exd.
12368
Murray, J. C. (1994). Keloids and hypertrophic scars. Clin. Dermatol. 12, 27–37.
Murray, J. C., Pollack, S. V., and Pinnell, S. R. (1981). Keloids: a review. J. Am.
Acad. Dermatol. 4, 461–470.
Naitoh, M., Hosokawa, N., Kubota, H., Tanaka, T., Shirane, H., Sawada, M., et al.
(2001). Upregulation of HSP47 and collagen type III in the dermal fibrotic
disease, keloid. Biochem. Biophys. Res. Commun. 280, 1316–1322. doi: 10.1006/
bbrc.2001.4257
Nakashima, M., Chung, S., Takahashi, A., Kamatani, N., Kawaguchi, T., Tsunoda,
T., et al. (2010). A genome-wide association study identifies four susceptibility
loci for keloid in the Japanese population. Nat. Genet. 42, 780–783. doi: 10.1038/
ng.645
Naylor, M. C., Lazar, D. A., Zamora, I. J., Mushin, O. P., Yu, L., Brissett, A. E., et al.
(2012). Increased in vitro differentiation of fibrocytes from keloid patients is
inhibited by serum amyloid P. Wound Repair Regen. 20, 277–283. doi: 10.1111/
j.1524-475X.2012.00782.x
Nemeth, A. J. (1993). Keloids and hypertrophic scars. J. Dermatol. Surg. Oncol. 19,
738–746.
Niessen, F. B., Spauwen, P. H., Schalkwijk, J., and Kon, M. (1999). On the nature of
hypertrophic scars and keloids: a review. Plast. Reconstr. Surg. 104, 1435–1458.
doi: 10.1097/00006534-199910000- 00031
Ogawa, R. (2011). Mechanobiology of scarring. Wound Repair Regen. 19(Suppl. 1),
s2–s9. doi: 10.1111/j.1524-475X.2011.00707.x
Ogawa, R. (2017). Keloid and hypertrophic scars are the result of chronic
inflammation in the reticular dermis. Int. J. Mol. Sci. 18, 1–10. doi: 10.3390/
ijms18030606
Ogawa, R., and Akaishi, S. (2016). Endothelial dysfunction may play a key role in
keloid and hypertrophic scar pathogenesis keloids and hypertrophic scars may
be vascular disorders. Med. Hypotheses 96, 51–60. doi: 10.1016/j.mehy.2016.09.
024
Ogawa, R., Akaishi, S., and Izumi, M. (2009). Histologic analysis of keloids
and hypertrophic scars. Ann. Plast. Surg. 62, 104–105. doi: 10.1097/SAP.
0b013e3181855172
Ogawa, R., Arima, J., Ono, S., and Hyakusoku, H. (2013). Total management
of a severe case of systemic keloids associated with high blood pressure
(hypertension): clinical symptoms of keloids may be aggravated by
hypertension. Eplasty 13, 184–190.
Ogawa, R., Mitsuhashi, K., Hyakusoku, H., and Miyashita, T. (2003). Postoperative
electron-beam irradiation therapy for keloids and hypertrophic: retrospective
study of 147 cases followed for more than 18 months. Plast. Reconstr. Surg. 111,
547–553. doi: 10.1097/01.PRS.0000040466.55214.35
Ogawa, R., Okai, K., Tokumura, F., Mori, K., Ohmori, Y., Huang, C., et al. (2012).
The relationship between skin stretching/contraction and pathologic scarring:
the important role of mechanical forces in keloid generation. Wound Repair
Regen. 20, 149–157. doi: 10.1111/j.1524-475X.2012.00766.x
Ohtsuru, A., Yoshimoto, H., Ishihara, H., Namba, H., and Yamashita, S. (2000).
Insulin-like growth factor-I (IGF-I)/IGF-I receptor axis and increased invasion
activity of fibroblasts in keloid. Endocr. J. 47, S41–S44. doi: 10.1507/endocrj.47.
supplmarch_s41
Oluwasanmi, J. O. (1974). Keloids in the African. Clin Plast Surg. 1, 179–195.
Ong, C. T., Khoo, Y. T., Mukhopadhyay, A., Do, D. V., Lim, I. J., Aalami, O., et al.
(2007a). mTOR as a potential therapeutic target for treatment of keloids and
excessive scars. Exp. Dermatol. 16, 394–404. doi: 10.1111/j.1600-0625.2007.
00550.x
Ong, C. T., Khoo, Y. T., Mukhopadhyay, A., Masilamani, J., Do, D. V., Lim, I. J.,
et al. (2010). Comparative proteomic analysis between normal skin and keloid
scar. Br. J. Dermatol. 162, 1302–1315. doi: 10.1111/j.1365-2133.2010.09660.x
Ong, C. T., Khoo, Y. T., Tan, E. K., Mukhopadhyay, A., Do, D. V., Han, H. C., et al.
(2007b). Epithelial–mesenchymal interactions in keloid pathogenesis modulate
vascular endothelial growth factor expression and secretion. J. Pathol. 211,
95–108. doi: 10.1002/path.2081
Ooi, B. N. S., Mukhopadhyay, A., Masilamani, J., Do, D. V., Lim, C. P., Cao,
X. M., et al. (2010). Hepatoma-derived growth factor and its role in keloid
pathogenesis. J. Cell. Mol. Med. 14, 1328–1337. doi: 10.1111/j.1582-4934.2009.
00779.x
Ozawa, T., Okamura, T., Harada, T., Muraoka, M., Ozawa, N., Koyama, K., et al.
(2006). Accumulation of glucose in keloids with FDG-PET. Ann. Nucl. Med. 20,
41–44. doi: 10.1007/bf02985589
Park, G., Yoon, B. S., Moon, J. H., Kim, B., Jun, E. K., Oh, S., et al.
(2008). Green tea polyphenol epigallocatechin-3-gallate suppresses collagen
production and proliferation in keloid fibroblasts via inhibition of the STAT3-
signaling pathway. J. Invest. Dermatol. 128, 2429–2441. doi: 10.1038/jid.
2008.103
Peacock, E. E., Madden, J. W., and Trier, W. C. (1970). Biological basis for the
treatment of keloids and hypertrophic scars. South Med. J. 63, 755–760. doi:
10.1097/00007611-197007000- 00002
Perez, R., and Davis, S. C. (2008). Relevance of animal models for wound healing.
Wounds 20, 3–8.
Phan, T. T., Lim, I. J., Bay, B. H., Qi, R., Huynh, H. T., Lee, S. T., et al.
(2002). Differences in collagen production between normal and keloid-derived
Frontiers in Cell and Developmental Biology | www.frontiersin.org 23 May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 24
Limandjaja et al. The Keloid Disorder
fibroblasts in serum-media co-culture with keloid-derived keratinocytes.
J. Dermatol. Sci. 29, 26–34. doi: 10.1016/s0923- 1811(02)00008-7
Phan, T. T., Lim, I. J., Bay, B. H., Qi, R., Longaker, M. T., Lee, S. T., et al. (2003). Role
of IGF system of mitogens in the induction of fibroblast proliferation by keloid-
derived keratinocytes in vitro. Am. J. Physiol. Cell Physiol. 284, C860–C869.
doi: 10.1152/ajpcell.00350.2002
Plikus, M. V., Guerrero-Juarez, C. F., Ito, M., Li, Y. R., Dedhia, P. H., Zheng,
Y., et al. (2017). Regeneration of fat cells from myofibroblasts during wound
healing. Science 355, 748–752. doi: 10.1126/science.aai8792
Prathiba, V., Rao, K. S., and Gupta, P. D. (2001). Altered expression of keratins
during abnormal wound healing in human skin. Cytobios 104, 43–51.
Qu, M., Song, N., Chai, G., Wu, X., and Liu, W. (2013). Pathological niche
environment transforms dermal stem cells to keloid stem cells: a hypothesis of
keloid formation and development. Med. Hypotheses 81, 807–812. doi: 10.1016/
j.mehy.2013.08.033
Quan, T. E., Cowper, S., Wu, S. P., Bockenstedt, L. K., and Bucala, R. (2004).
Circulating fibrocytes: collagen-secreting cells of the peripheral blood. Int. J.
Biochem. Cell Biol. 36, 598–606. doi: 10.1016/j.biocel.2003.10.005
Ramos, M. L. C., Gragnani, A., and Ferreira, L. M. (2008). Is there an ideal
animal model to study hypertrophic scarring? J. Burn Care Res. 29, 363–368.
doi: 10.1097/BCR.0b013e3181667557
Robles, D., and Berg, D. (2007). Abnormal wound healing: keloids. Clin. Dermatol.
25, 26–32.
Rockwell, W. B., Cohen, I. K., and Ehrlich, H. P. (1998). Keloids and hypertrophic
scars: a comprehensive review. Plast. Reconstr. Surg. 84, 827–837. doi: 10.1097/
00006534-198911000- 00021
Romero-Valdovinos, M., Cárdenas-Mejía, A., Gutiérrez-Gómez, C., Flisser, A.,
Kawa-Karasik, S., and Ortiz-Monasterio, F. (2011). Keloid skin scars: the
influence of hyperbaric oxygenation on fibroblast growth and on the expression
of messenger RNA for insulin like growth factor and for transforming growth
factor. Vitr.Cell. De v. Biol.Anim. 47, 421–424. doi: 10.1007/s11626- 011-9418- 3
Russell, S. B., Russell, J. D., Trupin, K. M., Gayden, A. E., Opalenik, S. R., Nanney,
L. B., et al. (2010). Epigenetically altered wound healing in keloid fibroblasts.
J. Invest. Dermatol. 130, 2489–2496. doi: 10.1038/jid.2010.162
Russell, S. B., Trupin, J. S., Kennedy, R. Z., Russell, J. D., and Davidson, J. M. (1995).
Glucocorticoid regulation of elastin synthesis in human fibroblasts: down-
regulation in fibroblasts from normal dermis but not from keloids. J. Invest.
Dermatol. 104, 241–245. doi: 10.1111/1523- 1747.ep12612788
Russell, S. B., Trupin, J. S., Myers, J. C., Broquist, A. H., Smith, J. C., Myles,
M. E., et al. (1989). Differential glucocorticoid regulation of collagen mRNAs
in human dermal fibroblasts. Keloid-derived and fetal fibroblasts are refractory
to down-regulation. J. Biol. Chem. 264, 13730–13735.
Russell, S. B., Trupin, K. M., Rodríguez-Eaton, S., Russell, J. D., and Trupin, J. S.
(1988). Reduced growth-factor requirement of keloid-derived fibroblasts may
account for tumor growth. Proc. Natl. Acad. Sci. U.S.A. 85, 587–591. doi:
10.1073/pnas.85.2.587
Rutherford, A., and Glass, D. A. II (2017). A case–control study analyzing the
association of keloids with hypertension and obesity. Int. J. Dermatol. 56,
e176–e191. doi: 10.1111/ijd.13618
Saffari, T. M., Bijlard, E., van Bodegraven, E. A. M., Mureau, M. A. M., Hovius,
S. E. R., and Huygen, F. J. P. M. (2018). Sensory perception and nerve
fibre innervation in patients with keloid scars: an investigative study. Eur. J.
Dermatol. 28, 828–829. doi: 10.1684/ejd.2018.3405
Sano, H., Hokazono, Y., and Ogawa, R. (2018). Distensibility and gross elasticity of
the skin at various body sites and association with pathological scarring: a case
study. J. Clin. Aesthet. Dermatol. 11, 15–18.
Santucci, M., Borgognoni, L., Reali, U. M., and Gabbiani, G. (2001). Keloids
and hypertrophic scars of Caucasians show distinctive morphologic and
immunophenotypic profiles. Virchows Arch. 438, 457–463. doi: 10.1007/
s004280000335
Sato, M., Ishikawa, O., and Miyachi, Y. (1998). Distinct patterns of collagen
gene expression are seen in normal and keloid fibroblasts grown in three-
dimensional culture. Br. J. Dermatol. 138, 938–943. doi: 10.1046/j.1365- 2133.
1998.02258.x
Sayah, D. N., Soo, C., Shaw, W. W., Watson, J., Messadi, D., Longaker, M. T., et al.
(1999). Downregulation of apoptosis-related genes in keloid tissues. J. Surg. Res.
87, 209–216. doi: 10.1006/jsre.1999.5761
Seifert, O., Bayat, A., Geffers, R., Dienus, K., Buer, J., Löfgren, S., et al. (2008).
Identification of unique gene expression patterns within different lesional sites
of keloids. Wound Repair Regen. 16, 254–265. doi: 10.1111/j.1524-475X.2007.
00343.x
Seifert, O., and Mrowietz, U. (2009). Keloid scarring: bench and bedside. Arch.
Dermatol. Res. 301, 259–272. doi: 10.1007/s00403-009-0952-8
Seo, B. F., Lee, J. Y., and Jung, S.-N. (2013). Models of abnormal scarring. Biomed.
Res. Int. 2013:423147. doi: 10.1155/2013/423147
Seok, J., Warren, H. S., Cuenca, A. G., Mindrinos, M. N., Baker, H. V., Xu,
W., et al. (2013). Genomic responses in mouse models poorly mimic human
inflammatory diseases. Proc. Natl. Acad. Sci. U.S.A. 110, 3507–3512. doi: 10.
1073/pnas.1222878110
Shaker, S. A., Ayuob, N. N., and Hajrah, N. H. (2011). Cell talk: a
phenomenon observed in the keloid scar by immunohistochemical study.
Appl. Immunohistochem. Mol. Morphol. 19, 153–159. doi: 10.1097/PAI.
0b013e3181efa2ef
Shang, T., Yao, B., Gao, D., Xie, J., Fu, X., and Huang, S. (2018). A novel model of
humanised keloid scarring in mice. Int. Wound J. 15, 90–94. doi: 10.1111/iwj.
12838
Sharquie, K. E., and Al-Dhalimi, M. A. (2003). Keloid in Iraqi patients: a
clinicohistopathologic study. Dermatol. Surg. 29, 847–851. doi: 10.1046/j.1524-
4725.2003.29230.x
Shih, B., and Bayat, A. (2010). Genetics of keloid scarring. Arch. Dermatol. Res. 302,
319–339. doi: 10.1007/s00403-009- 1014-y
Shih, B., Garside, E., McGrouther, D. A., and Bayat, A. (2010). Molecular dissection
of abnormal wound healing processes resulting in keloid disease. Wound Repair
Regen. 18, 139–153. doi: 10.1111/j.1524-475X.2009.00553.x
Shin, J. U., Kim, S. H., Kim, H., Noh, J. Y., Jin, S., Park, C. O., et al. (2016). TSLP
is a potential initiator of collagen synthesis and an activator of CXCR4/SDF-1
axis in keloid pathogenesis. J. Invest. Dermatol. 136, 507–515. doi: 10.1016/j.jid.
2015.11.008
Sible, J. C., Eriksson, E., Smith, S. P., and Oliver, N. (1994). Fibronectin gene
expression differs in normal and abnormal human wound healing. Wound
Repair Regen. 2, 3–19. doi: 10.1046/j.1524- 475X.1994.20104.x
Sidgwick, G. P., Iqbal, S. A., and Bayat, A. (2013). Altered expression of hyaluronan
synthase and hyaluronidase mRNA may affect hyaluronic acid distribution
in keloid disease compared with normal skin. Exp. Dermatol. 22, 377–379.
doi: 10.1111/exd.12147
Slemp, A. E., and Kirschner, R. E. (2006). Keloids and scars: a review of keloids and
scars, their pathogenesis, risk factors, and management. Curr. Opin. Pediatr. 18,
396–402. doi: 10.1097/01.mop.0000236389.41462.ef
Slominski, A., Wortsman, J., Mazurkiewicz, J. E., Matsuoka, L., Dietrich, J.,
Lawrence, K., et al. (1993). Detection of proopiomelanocortin-derived antigens
in normal and pathologic human skin. J. Lab. Clin. Med. 122, 658–666.
Smith, J. C., Boone, B. E., Opalenik, S. R., Williams, S. M., and Russell, S. B.
(2008). Gene profiling of keloid fibroblasts shows altered expression in multiple
fibrosis-associated pathways. J. Invest. Dermatol. 128, 1298–1310. doi: 10.1038/
sj.jid.5701149
Snyder, A. L., Zmuda, J. M., and Thompson, P. D. (1996). Keloid associated with
hypertension. Lancet 347, 465–466.
Sogabe, Y., Akimoto, S., Abe, M., Ishikawa, O., Takagi, Y., and Imokawa, G.
(2002). Functions of the stratum corneum in systemic sclerosis as distinct
from hypertrophic scar and keloid functions. J. Dermatol. Sci. 29, 49–53. doi:
10.1016/s0923-1811(02)00006- 3
Song, C. (2014). Hypertrophic scars and keloids in surgery: current concepts. Ann.
Plast. Surg. 73(Suppl. 1), S108–S118. doi: 10.1097/SAP.0000000000000256
Stanisz, H., Seifert, M., Tilgen, W., Vogt, T., and Rass, K. (2011). Reciprocal
responses of fibroblasts and melanocytes to α-MSH depending on MC1R
polymorphisms. Dermatoendocrinol 3, 259–265. doi: 10.4161/derm.3.4.
17454
Stewart, J., and Glass, I. I. D. A. (2018). Plasma angiotensin-converting enzyme
levels in patients with keloids and/or hypertension. Wounds 30, E71–E72.
Stone, R. C., Pastar, I., Ojeh, N., Chen, V., Liu, S., Garzon, K. I., et al. (2017).
Epithelial-mesenchymal transition in tissue repair and fibrosis. Cell Tissue Res.
365, 495–506. doi: 10.1007/s00441- 016-2464-0
Suarez, E., Syed, F., Alonso-Rasgado, T., and Bayat, A. (2015). Identification of
biomarkers involved in differential profiling of hypertrophic and keloid scars
Frontiers in Cell and Developmental Biology | www.frontiersin.org 24 May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 25
Limandjaja et al. The Keloid Disorder
versus normal skin. Arch. Dermatol. Res. 307, 115–133. doi: 10.1007/s00403-
014-1512- 4
Suarez, E., Syed, F., Alonso-Rasgado, T., Mandal, P., and Bayat, A. (2013).
Up-regulation of tension-related proteins in keloids: knockdown of HSP27,
α2β1-integrin, and PAI-2 shows convincing reduction of extracellular
matrix production. Plast. Reconstr. Surg. 131, 158–173. doi: 10.1097/PRS.
0b013e3182789b2b
Suetake, T., Sasai, S., Zhen, Y. X., Ohi, T., and Tagami, H. (1996). Functional
analyses of the stratum corneum in scars. Sequential studies after injury
and comparison among keloids, hypertrophic scars, and atrophic scars. Arch.
Dermatol. 132, 1453–1458.
Sun, L.-M., Wang, K.-H., and Lee, Y.-C. G. (2014). Keloid incidence in Asian
people and its comorbidity with other fibrosis-related diseases: a nationwide
population-based study. Arch. Dermatol. Res. 306, 803–808. doi: 10.1007/
s00403-014- 1491-5
Supp, D. M., Glaser, K., Hahn, J. M., McFarland, K. L., and Boyce, S. T. (2012a).
Abnormal responses of keloid tissue to wounding identified using in vitro
model system. Eplasty 12:e19.
Supp, D. M., Hahn, J. M., Glaser, K., McFarland, K. L., and Boyce, S. T.
(2012b). Deep and superficial keloid fibroblasts contribute differentially to
tissue phenotype in a novel in vivo model of keloid scar. Plast. Reconstr. Surg.
129, 1259–1271. doi: 10.1097/PRS.0b013e31824ecaa9
Supp, D. M., Hahn, J. M., McFarland, K. L., and Glaser, K. (2014).
Inhibition of hyaluronan synthase 2 reduces the abnormal migration rate
of keloid keratinocytes. J. Burn Care Res. 35, 84–92. doi: 10.1097/BCR.
0b013e3182a2a9dd
Suttho, D., Mankhetkorn, S., Binda, D., Pazart, L., Humbert, P., and Rolin, G.
(2017). 3D modeling of keloid scars in vitro by cell and tissue engineering. Arch.
Dermatol. Res. 309, 55–62. doi: 10.1007/s00403-016-1703-2
Suzawa, H., Kikuchi, S., Arai, N., and Koda, A. (1992). The mechanism involved in
the inhibitory action of tranilast on collagen biosynthesis of keloid fibroblasts.
Jpn. J. Pharmacol. 60, 91–96. doi: 10.1254/jjp.60.91
Syed, F., Ahmadi, E., Iqbal, S. A., Singh, S., McGrouther, D. A., and Bayat, A.
(2011). Fibroblasts from the growing margin of keloid scars produce higher
levels of collagen I and III compared with intralesional and extralesional sites:
clinical implications for lesional site-directed therapy. Br. J. Dermatol. 164,
83–96. doi: 10.1111/j.1365-2133.2010.10048.x
Syed, F., Bagabir, R. A., Paus, R., and Bayat, A. (2013a). Ex vivo evaluation
of antifibrotic compounds in skin scarring: EGCG and silencing of PAI-1
independently inhibit growth and induce keloid shrinkage. Lab. Investig. 93,
946–960. doi: 10.1038/labinvest.2013.82
Syed, F., and Bayat, A. (2012). Notch signaling pathway in keloid disease: enhanced
fibroblast activity in a Jagged-1 peptide-dependent manner in lesional vs.
extralesional fibroblasts. Wound Repair Regen. 20, 688–706. doi: 10.1111/j.
1524-475X.2012.00823.x
Syed, F., Sanganee, H. J., Bahl, A., and Bayat, A. (2013b). Potent dual inhibitors of
TORC1 and TORC2 complexes (KU-0063794 and KU-0068650) demonstrate
in vitro and ex vivo anti-keloid scar activity. J. Invest. Dermatol. 133, 1340–1350.
doi: 10.1038/jid.2012.483
Syed, F., Sherris, D., Paus, R., Varmeh, S., and Pandolfi, P. P. (2012). Keloid disease
can be inhibited by antagonizing excessive mTOR signaling with a novel dual
TORC1/2 inhibitor. Am. J. Pathol. 181, 1642–1658. doi: 10.1016/j.ajpath.2012.
08.006
Szulgit, G., Rudolph, R., Wandel, A., Tenenhaus, M., Panos, R., and Gardner, H.
(2002). Alterations in fibroblast α1β1 integrin collagen receptor expression in
keloids and hypertrophic scars. J. Invest. Dermatol. 118, 409–415. doi: 10.1046/
j.0022-202x.2001.01680.x
Tan, K. T., McGrouther, D. A., Day, A. J., Milner, C. M., and Bayat, A.
(2011). Characterization of hyaluronan and TSG-6 in skin scarring: differential
distribution in keloid scars, normal scars and unscarred skin. J. Eur. Acad.
Dermatol. Venereol. 25, 317–327. doi: 10.1111/j.1468-3083.2010.03792.x
Tanaka, A., Hatoko, M., Tada, H., Iioka, H., Niitsuma, K., and Miyagawa, S. (2004).
Expression of p53 family in scars. J. Dermatol. Sci. 34, 17–24. doi: 10.1016/j.
jdermsci.2003.09.005
Teofoli, P., Motoki, K., Lotti, T. M., Uitto, J., and Mauviel, A. (1997).
Propiomelanocortin (POMC) gene expression by normal skin and keloid
fibroblasts in culture: modulation by cytokines. Exp. Dermatol. 6, 111–115.
doi: 10.1111/j.1600-0625.1997.tb00156.x
Theoret, C. L., Acvs, D., Olutoye, O. O., Parnell, L. K. S., and Hicks, J.
(2013). Equine exuberant granulation tissue and human keloids: a comparative
histopathologic study. Vet. Surg. 42, 783–789. doi: 10.1111/j.1532-950X.2013.
12055.x
Touchi, R., Ueda, K., Kurokawa, N., and Tsuji, M. (2016). Central regions of
keloids are severely ischaemic. J. Plast. Reconstr. Aesthetic Surg. 69, e35–e41.
doi: 10.1016/j.bjps.2015.11.006
Trusler, H. M., and Bauer, T. B. (1948). Keloids and hypertrophic scars. Arch. Surg.
57, 539–552.
Tsujita-Kyutoku, M., Uehara, N., Matsuoka, Y., Kyutoku, S., Ogawa, Y., and
Tsubura, A. (2005). Comparison of transforming growth factor-beta/Smad
signaling between normal dermal fibroblasts and fibroblasts derived from
central and peripheral areas of keloid lesions. In Vivo (Brooklyn) 19, 959–963.
Tuan, T. L., and Nichter, L. S. (1998). The molecular basis of keloid and
hypertrophic scar formation. Mol. Med. Today 4, 19–24. doi: 10.1016/S1357-
4310(97)80541-2
Tucci-Viegas, V. M., Hochman, B., Frana, J. P., and Ferreira, L. M. (2010). Keloid
explant culture: a model for keloid fibroblasts isolation and cultivation based
on the biological differences of its specific regions. Int. Wound J. 7, 339–348.
doi: 10.1111/j.1742-481X.2010.00698.x
Tulandi, T., Al-Sannan, B., Akbar, G., Ziegler, C., and Miner, L. (2011). Prospective
study of intraabdominal adhesions among women of different races with or
without keloids. Am. J. Obs. Gynecol. 204, e1–e4. doi: 10.1016/j.ajog.2010.
09.005
Uchida, G., Yoshimura, K., Kitano, Y., Okazaki, M., and Harii, K. (2003). Tretinoin
reverses upregulation of matrix metalloproteinase-13 in human keloid-derived
fibroblasts. Exp. Dermatol. 12(Suppl. 2), 35–42. doi: 10.1034/j.1600-0625.12.
s2.6.x
Ueda, K., Furuya, E., Yasuda, Y., Oba, S., and Tajima, S. (1999). Keloids have
continuous high metabolic activity. Plast. Reconstr. Surg. 104, 694–698. doi:
10.1097/00006534-199909030- 00012
Ueda, K., Yasuda, Y., Furuya, E., and Oba, S. (2004). Inadequate blood supply
persists in keloids. Scand. J. Plast. Reconstr. Surg. Hand Surg. 38, 267–271.
doi: 10.1080/02844310410029552
Uitto, J., Perejda, A. J., Abergel, R. P., Chu, M. L., and Ramirez, F. (1985). Altered
steady-state ratio of type I/III procollagen mRNAs correlates with selectively
increased type I procollagen biosynthesis in cultured keloid fibroblasts. Proc.
Natl. Acad. Sci. U.S.A. 82, 5935–5939. doi: 10.1073/pnas.82.17.5935
Uzawa, K., Yeowell, H. N., Yamamoto, K., Mochida, Y., Tanzawa, H., and
Yamauchi, M. (2003). Lysine hydroxylation of collagen in a fibroblast cell
culture system. Biochem. Biophys. Res. Commun. 305, 484–487. doi: 10.1016/
s0006-291x(03)00799- x
van der Slot, A. J., Zuurmond, A., van Den Bogaerdt, A. J., Ulrich, M. M. W.,
Middelkoop, E., Boers, W., et al. (2004). Increased formation of pyridinoline
cross-links due to higher telopeptide lysyl hydroxylase levels is a general fibrotic
phenomenon. Matrix Biol. 23, 251–257. doi: 10.1016/j.matbio.2004.06.001
Varmeh, S., Egia, A., McGrouther, D., Tahan, S. R., Bayat, A., and Pandolfi, P. P.
(2011). Cellular senescence as a possible mechanism for halting progression of
keloid lesions. Genes Cancer 2, 1061–1066. doi: 10.1177/1947601912440877
Velez Edwards, D. R., Tsosie, K., Williams, S. M., Edwards, T. L., and Russell,
S. B. (2014). Admixture mapping identifies a locus at 15q21.2-22.3 associated
with keloid formation in African Americans. Hum. Genet. 133, 1513–1523.
doi: 10.1007/s00439-014- 1490-9
Verhaegen, P. D. H. M., Van Zuijlen, P. P. M., Pennings, N. M., Van Marle, J.,
Niessen, F. B., Van Der Horst, C. M. A. M., et al. (2009). Differences in collagen
architecture between keloid, hypertrophic scar, normotrophic scar, and normal
skin: an objective histopathological analysis. Wound Repair Regen. 17, 649–656.
doi: 10.1111/j.1524-475X.2009.00533.x
Videira, I. F. S., Moura, D. F. L., and Magina, S. (2013). Mechanisms regulating
melanogenesis. An. Bras. Dermatol. 88, 76–83.
Vincent, A. S., Phan, T. T., Mukhopadhyay, A., Lim, H. Y., Halliwell, B., and Wong,
K. P. (2008). Human skin keloid fibroblasts display bioenergetics of cancer cells.
J. Invest. Dermatol. 128, 702–709. doi: 10.1038/sj.jid.5701107
Wadhwa, R., Priyandoko, D., Gao, R., Widodo, N., Nigam, N., and Kaul, S. C.
(2016). Stress chaperone mortalin regulates human melanogenesis. Cell Stress
Chaperones 21, 631–644. doi: 10.1007/s12192-016-0688-2
Wang, X., Liu, K., Ruan, M., Yang, J., and Gao, Z. (2018). Gallic acid inhibits
fibroblast growth and migration in keloids through the AKT/ERK signaling
Frontiers in Cell and Developmental Biology | www.frontiersin.org 25 May 2020 | Volume 8 | Article 360
fcell-08-00360 May 22, 2020 Time: 19:44 # 26
Limandjaja et al. The Keloid Disorder
pathway. Acta Biochim. Biophys. Sin. 50, 1114–1120. doi: 10.1093/abbs/
gmy115
Wang, X., Liu, Y., Chen, X., Zhang, M., and Xiao, Z. (2013). Impact of miR-21
on the expression of FasL in the presence of TGF-β1. Aesthetic Surg. J. 33,
1186–1198. doi: 10.1177/1090820X13511969
Wen, A., Chua, C., Uin, S., Ting, Y., Fu, Z., Kang, C., et al. (2011). Keloid fibroblasts
are more sensitive to Wnt3a treatment in terms of elevated cellular growth and
fibronectin expression. J. Dermatol. Sci. 64, 199–209. doi: 10.1016/j.jdermsci.
2011.09.008
Wirohadidjojo, Y. W., Radiono, S., Budiyanto, A., and Soebono, H. (2011). Cellular
viability, collagen deposition, and transforming growth factor β1 production
among ultraviolet B-irradiated keloid fibroblasts. Aesthetic Plast. Surg. 35,
1050–1055. doi: 10.1007/s00266- 011-9732-x
Witt, E., Maliri, A., McGrouther, D. A., and Bayat, A. (2008). RAC activity in
keloid disease: comparative analysis of fibroblasts from margin of keloid to its
surrounding normal skin. Eplasty 8:e19.
Wolfram, D., Tzankov, A., Pülzl, P., and Piza-Katzer, H. (2009). Hypertrophic scars
and keloids a review of their pathophysiology, risk factors, and therapeutic
management. Dermatologic Surg. 35, 171–181. doi: 10.1111/j.1524- 4725.2008.
34406.x
Wong, T.-W., and Lee, J. Y.-Y. (2008). Should excised keloid scars be sent
for routine histologic analysis? Ann. Plast. Surg. 60:724. doi: 10.1097/SAP.
0b013e318178d9f1
Wu, Y., Zhang, Q., Ann, D. K., Akhondzadeh, A., Duong, H. S., Messadi, D. V.,
et al. (2004). Increased vascular endothelial growth factor may account for
elevated level of plasminogen activator inhibitor-1 via activating ERK1/2 in
keloid fibroblasts. Am. J. Physiol. Cell Physiol. 286, C905–C912. doi: 10.1152/
ajpcell.00200.2003
Xia, W., Kong, W., Wang, Z., Phan, T. T., Lim, I. J., Longaker, M. T., et al. (2007).
Increased CCN2 transcription in keloid fibroblasts requires cooperativity
between AP-1 and SMAD binding sites. Ann. Surg. 246, 886–895. doi: 10.1097/
SLA.0b013e318070d54f
Xia, W., Longaker, M. T., and Yang, G. P. (2005). P38 MAP kinase mediates
transforming growth factor-β2 transcription in human keloid fibroblasts. Am.
J. Physiol. Regul. Integr. Comp. Physiol. 290, R501–R508. doi: 10.1152/ajpregu.
00472.2005
Xia, W., Phan, T.-T., Lim, I. J., Longaker, M. T., and Yang, G. P. (2004). Complex
epithelial-mesenchymal interactions modulate transforming growth factor-beta
expression in keloid-derived cells. Wound Repair Regen. 12, 546–556. doi:
10.1111/j.1067-1927.2004.012507.x
Xin, Y., Wang, X., Zhu, M., Qu, M., Bogari, M., Lin, L., et al. (2017). Expansion of
CD26 positive fibroblast population promotes keloid progression. Exp. Cell Res.
356, 104–113. doi: 10.1016/j.yexcr.2017.04.021
Xue, H., McCauley, R. L., and Zhang, W. (2000). Elevated interleukin-6
expression in keloid fibroblasts. J. Surg. Res. 89, 74–77. doi: 10.1006/jsre.1999.
5805
Yadav, T. P., Mishra, S., and Gautam, R. K. (1995). Keloid at the venipuncture site.
Indian Pediatr. 32, 1243–1244.
Yagi, K. I., Dafalla, A. A., and Osman, A. A. (1979). Does an immune reaction to
sebum in wounds cause keloid scars? Beneficial effects of desensitisation. Br. J.
Plast. Surg. 32, 223–225. doi: 10.1016/s0007- 1226(79)90037-7
Yagi, Y., Muroga, E., Naitoh, M., Isogai, Z., Matsui, S., Ikehara, S., et al. (2013).
An ex vivo model employing keloid-derived cell-seeded collagen sponges for
therapy development. J. Invest. Dermatol. 133, 386–393. doi: 10.1038/jid.2012.
314
Yan, C., Grimm, W. A., Garner, W. L., Qin, L., Travis, T., Tan, N., et al. (2010).
Epithelial to mesenchymal transition in human skin wound healing is induced
by tumor necrosis factor-αthrough bone morphogenic protein-2. Am. J. Pathol.
176, 2247–2258. doi: 10.2353/ajpath.2010.090048
Yan, L., Cao, R., Wang, L., Liu, Y., Pan, B., Yin, Y., et al. (2015). Epithelial-
mesenchymal transition in keloid tissues and TGF-β1-induced hair follicle
outer root sheath keratinocytes. Wound Repair Regen. 23, 601–610. doi: 10.
1111/wrr.12320
Yang, E., Qipa, Z., and Hengshu, Z. (2014). The expression of DNMT1 in
pathologic scar fibroblasts and the effect of 5-aza-2-deoxycytidine on cytokines
of pathologic scar fibroblasts. Wounds 26, 139–146.
Yang, G., Lim, I., Phan, T., Lorenz, H., and Longaker, M. (2003). From scarless fetal
wounds to keloids: molecular studies in wound healing. Wound Repair Regen.
11, 411–418. doi: 10.1046/j.1524- 475x.2003.11604.x
Yeh, F., Shen, H., Lin, M., Chang, C., Tai, H.-Y., and Huang, M.-H. (2006). Keloid-
derived fibroblasts have a diminished capacity to produce prostaglandin E2.
Burns 32, 299–304. doi: 10.1016/j.burns.2005.10.009
Yeh, F. L., Shen, H.-D., and Tai, H.-Y. (2009). Decreased production of MCP-1 and
MMP-2 by keloid-derived fibroblasts. Burns 35, 348–351. doi: 10.1016/j.burns.
2008.06.018
Yoshimoto, H., Ishihara, H., Ohtsuru, A., Akino, K., Murakami, R., Kuroda, H.,
et al. (1999). Overexpression of insulin-like growth factor-1 (IGF-I) receptor
and the invasiveness of cultured keloid fibroblasts. Am. J. Pathol. Am. Soc.
Invest. Pathol. 154, 883–889. doi: 10.1016/S0002-9440(10)65335- 7
Young, W. G., Worsham, M. J., Joseph, C. L. M., Divine, G. W., and Jones, L. R. D.
(2014). Incidence of keloid and risk factors following head and neck surgery.
JAMA Facial Plast. Surg. 16, 379–380. doi: 10.1001/jamafacial.2014.113
Yu, D., Shang, Y., Yuan, J., Ding, S., Luo, S., and Hao, L. (2016). Wnt/β-catenin
signaling exacerbates keloid cell proliferation by regulating telomerase. Cell.
Physiol. Biochem. 39, 2001–2013. doi: 10.1159/000447896
Zhang, G., Gao, W., Li, X., Yi, C., Zheng, Y., Li, Y., et al. (2009). Effect of
camptothecin on collagen synthesis in fibroblasts from patients with keloid.
Ann. Plast. Surg. 63, 94–99. doi: 10.1097/SAP.0b013e3181872775
Zhang, Q., Kelly, A. P., Wang, L., French, S. W., Tang,X., Duong, H. S., et al. (2006).
Green tea extract and (-)-epigallocatechin-3-gallate inhibit mast cell-stimulated
type I collagen expression in keloid fibroblasts via blocking PI-3K/Akt signaling
pathways. J. Invest. Dermatol. 126, 2607–2613. doi: 10.1038/sj.jid.5700472
Zhang, Q., Yamaza, T., Kelly, A. P., Shi, S., Wang, S., Brown, J., et al. (2009). Tumor-
like stem cells derived from human keloid are governed by the inflammatory
niche driven by IL-17/IL-6 axis. PLoS One 4:e7798. doi: 10.1371/journal.pone.
0007798
Zhao, B., Guan, H., Liu, J. Q., Zheng, Z., Zhou, Q., Zhang, J., et al. (2017).
Hypoxia drives the transition of human dermal fibroblasts to a myofibroblast-
like phenotype via the TGF-β1/Smad3 pathway. Int. J. Mol. Med. 39, 153–159.
doi: 10.3892/ijmm.2019.4403
Zhou, H. M., Wang, J., Elliott, C., Wen, W., Hamilton, D. W., and Conway, S. J.
(2010). Spatiotemporal expression of periostin during skin development and
incisional wound healing: lessons for human fibrotic scar formation. J. Cell
Commun. Signal. 4, 99–107. doi: 10.1007/s12079- 010-0090-2
Zhu, F., Wu, B., Li, P., Wang, J., Tang, H., Liu, Y., et al. (2013). Association study
confirmed susceptibility loci with keloid in the Chinese Han population. PLoS
One 8:e62377. doi: 10.1371/journal.pone.0062377
Conflict of Interest: The authors declare that the research was conducted in the
absence of any commercial or financial relationships that could be construed as a
potential conflict of interest.
Copyright © 2020 Limandjaja, Niessen, Scheper and Gibbs. This is an open-access
article distributed under the terms of the Creative Commons Attribution License
(CC BY). The use, distribution or reproduction in other forums is permitted, provided
the original author(s) and the copyright owner(s) are credited and that the original
publication in this journal is cited, in accordance with accepted academicpractice. No
use, distribution or reproduction is permitted which does not comply with theseterms.
Frontiers in Cell and Developmental Biology | www.frontiersin.org 26 May 2020 | Volume 8 | Article 360
... Keloids are fibrous proliferative, prominent skin scar-like lesions of the skin, characterized by a persistent, gradual growth beyond the wound margins into the surrounding healthy skin. Due to the characteristics of keloids, they can cause severe pain, chronic pruritus, psychosocial impairment, and motor striction to the patients, which leaves a heavy burden [88][89][90][91][92]. ...
... Since keloids are mainly characterized by increased proliferation of fibroblasts, anextensive overproduction of ECM components, and the consequent decrease in MMPs, research on keloids has primarily focused on the role of fibroblasts in keloid developments [88][89][90][91][92]. However, recently, researchers have noticed that the surrounding immune microenvironment can play a vital role in the development of keloids. ...
... However, recently, researchers have noticed that the surrounding immune microenvironment can play a vital role in the development of keloids. The increased number of T cells, LCs, MCs, and macrophages in keloids as opposed to normal tissues is associated with skin fibrosis [88][89][90][91][92]. The keloid fibroblasts and immune cells can interact with each other and develop a synergistic and complex regulation of molecules and pathways, which contribute to the development of keloids. ...
Article
Full-text available
This review was written with the aim of examining the effects that cause an insult, such as a wound, to an organ, such as the skin. Before examining the cellular mechanisms relating to wound healing, the reader is invited to read about the structure of the skin as a necessary basis for understanding the final aim of this review. The structure of the skin as a basis for understanding the phenomena relating to wound healing is addressed, taking into account the updated literature that addresses the numerous problems of the skin microenvironment. Starting from this awareness, the paragraphs dedicated to wound healing become complicated when this phenomenon is not implemented and therefore while the problems of chronic wounds, keloids, and hypertrophic scars are addressed, these are pathologies that are still difficult to understand and treat today.
... The raised scars of KD extend beyond the boundaries of the original wounds, but an HTS is contained within the boundaries of the original injury [13][14][15]. KDs are preferentially induced at sites with high skin tension, such as the chest, shoulder, upper back, posterior neck, cheeks, knees, and earlobes, but HTSs can be induced anywhere, and there are no predominant anatomical sites [16,17]. The occurrence of KDs and HTSs is mostly concentrated in individuals 9-20 years old (HTS: 77.2%; KD: 81.8%) in China; the male sex is a risk factor for HTSs (adjusted p < 0.001), and KDs are associated with family history (adjusted p < 0.050). ...
... KDs do not occur in patients with albinism, suggesting that melanocytes play a possible role in KD formation. In addition, the contribution of different alleles of human leukocyte antigen is also reported in KD formation [17][18][19][20]. Differences in the structures of the basement membrane zone, such as the collagen structure in the dermal layers, and the frequency of mast cells between HD, HTSs and normal skin are also reported [21]. ...
... No predominant anatomical sites of HTSs. [16,17] Incidence Less common. Association with age ≥ 22 years, family history and genetic factors. ...
Article
Full-text available
Keloids (KD) and hypertrophic scars (HTS), which are quite raised and pigmented and have increased vascularization and cellularity, are formed due to the impaired healing process of cutaneous injuries in some individuals having family history and genetic factors. These scars decrease the quality of life (QOL) of patients greatly, due to the pain, itching, contracture, cosmetic problems, and so on, depending on the location of the scars. Treatment/prevention that will satisfy patients’ QOL is still under development. In this article, we review pharmacotherapy for treating KD and HTS, including the prevention of postsurgical recurrence (especially KD). Pharmacotherapy involves monotherapy using a single drug and combination pharmacotherapy using multiple drugs, where drugs are administered orally, topically and/or through intralesional injection. In addition, pharmacotherapy for KD/HTS is sometimes combined with surgical excision and/or with physical therapy such as cryotherapy, laser therapy, radiotherapy including brachytherapy, and silicone gel/sheeting. The results regarding the clinical effectiveness of each mono-pharmacotherapy for KD/HTS are not always consistent but rather scattered among researchers. Multimodal combination pharmacotherapy that targets multiple sites simultaneously is more effective than mono-pharmacotherapy. The literature was searched using PubMed, Google Scholar, and Online search engines.
... Cutaneous scars present patients with substantial hurdles because they may cause serious health issues, such as aesthetic concerns, as well as complications like thickness, inflammation, and itching. Scarring has a complicated pathogenesis that includes, among other things, immunological alterations, stress in the wound, genetic variables, and programmed cell death [162]. Today, a number of methods are employed to prevent and treat elevated scars, including the use of pressure garments, silicone gels, corticosteroids, lasers, surgical treatments, etc. Pressure garments limit blood supply to the wound site, which lower oxygen content, enhance collagenase activity, and weaken collagen fiber attachment. ...
Article
The skin, serving as the body's outermost layer, boasts a vast area and intricate structure, functioning as the primary barrier against external threats. Disruptions; in the composition and functionality of...
... Under normal circumstances, these processes are well-coordinated, taking place and resolving promptly to restore skin homeostasis (1, 2). However, excessive mechanical stretching at the site of a skin wound may interfere with the course of wound healing, leading to the formation of enlarged scars (3), and hypertrophic scars or keloids in predisposed individuals (4)(5)(6). Keloids, typically triggered after skin injury, are bulging scar tissues that extend beyond the boundary of the original wound, composed mainly of hyper-proliferative fibroblasts, excessive collagen deposited by fibroblasts, and neo-vasculature. Body areas subject to heightened mechanical stretching, such as the anterior chest and shoulders, are prone to keloid formation (7). ...
Preprint
The healing of human skin wounds is susceptible to perturbation caused by excessive mechanical stretching, resulting in enlarged scars, hypertrophic scars, or even keloids in predisposed individuals. Keloids are fibro-proliferative scar tissues that extend beyond the initial wound boundary, consisting of the actively progressing leading edge and the quiescent center. The stretch-associated outgrowth and enhanced angiogenesis are two features of the leading edge of keloids. However, which cell population is responsible for transducing the mechanical stimulation to the pathological alterations of keloid tissues remains unclear. Herein, through joint analysis of single-cell RNA sequencing of keloid specimens and RNA sequencing of stretched keloid fibroblasts, we identified CD74 ⁺ fibroblasts, a previously unappreciated subset of fibroblasts, as a key player in stretch-induced keloid progression. Examination of macrophage markers suggested a possible myeloid origin of the CD74 ⁺ fibroblasts. Immunostaining of keloid cryosections depicted a predominant distribution of CD74 ⁺ fibroblasts in the leading edge, interacting with vasculature. CD74 ⁺ fibroblasts possessed pro-angiogenic and migratory capacities, as revealed by in vitro transwell and tube formation assays on purified CD74 ⁺ fibroblasts. Additionally, these cells underwent proliferation upon stretching, through PIEZO1-mediated calcium influx and the downstream ERK and AKT signaling. Collectively, our findings propose a model wherein CD74 ⁺ fibroblasts serve as pivotal drivers of stretch-induced keloid progression, fueled by their proliferative, pro-angiogenic, and migratory capacities. Targeting the attributes of CD74 ⁺ fibroblasts hold promise as a therapeutic strategy for keloid management. Significance statement Keloids are fibro-proliferative scars resulting from aberrant skin wound healing processes, consisting of the actively progressing leading edge and the quiescent center. Mechanical stretching and neo-vascularization have both been implicated in keloid progression, yet little is known about whether they are interconnected. Herein, we demonstrated that CD74 ⁺ fibroblasts, a previously undiscovered fibroblast subset, possessed heightened pro-angiogenic and migratory capacities, and underwent proliferation upon mechanical stretching, thereby facilitating the progression of the leading edge of keloids. Examination of macrophage markers suggested a possible myeloid origin of CD74 ⁺ fibroblasts. Our findings uncover the connection between stretch-induced keloid progression and neo-vascularization through CD74 ⁺ fibroblasts and provide valuable insights into potential therapeutic interventions.
Article
Background Keloid is a chronic proliferative fibrotic disease caused by abnormal fibroblasts proliferation and excessive extracellular matrix (ECM) production. Numerous fibrotic disorders are significantly influenced by ferroptosis, and targeting ferroptosis can effectively mitigate fibrosis development. This study aimed to investigate the role and mechanism of ferroptosis in keloid development. Methods Keloid tissues from keloid patients and normal skin tissues from healthy controls were collected. Iron content, lipid peroxidation (LPO) level, and the mRNA and protein expression of ferroptosis-related genes including solute carrier family 7 member 11 (SLC7A11), glutathione peroxidase 4 (GPX4), transferrin receptor (TFRC), and nuclear factor erythroid 2-related factor 2 (Nrf2) were determined. Mitochondrial morphology was observed using transmission electron microscopy (TEM). Keloid fibroblasts (KFs) were isolated from keloid tissues, and treated with ferroptosis inhibitor ferrostatin-1 (fer-1) or ferroptosis activator erastin. Iron content, ferroptosis-related marker levels, LPO level, mitochondrial membrane potential, ATP content, and mitochondrial morphology in KFs were detected. Furthermore, the protein levels of α-smooth muscle actin (α-SMA), collagen I, and collagen III were measured to investigate whether ferroptosis affect fibrosis in KFs. Results We found that iron content and LPO level were substantially elevated in keloid tissues and KFs. SLC7A11, GPX4, and Nrf2 were downregulated and TFRC was upregulated in keloid tissues and KFs. Mitochondria in keloid tissues and KFs exhibited ferroptosis-related pathology. Fer-1 treatment reduced iron content, restrained ferroptosis and mitochondrial dysfunction in KFs, Moreover, ferrostatin-1 restrained the protein expression of α-SMA, collagen I, and collagen III in KFs. Whereas erastin treatment showed the opposite results. Conclusion Ferroptosis exists in keloid. Ferrostatin-1 restrained ECM deposition and fibrosis in keloid through inhibiting ferroptosis, and erastin induced ECM deposition and fibrosis through intensifying ferroptosis.
Article
Recently, the pathomechanisms of keloids have been extensively researched using transcriptomic analysis, but most studies did not consider the activity of keloids. We aimed to profile the transcriptomics of keloids according to their clinical activity and location within the keloid lesion, compared with normal and mature scars. Tissue samples were collected (keloid based on its activity (active and inactive), mature scar from keloid patients and normal scar (NS) from non‐keloid patients). To reduce possible bias, all keloids assessed in this study had no treatment history and their location was limited to the upper chest or back. Multiomics assessment was performed by using single‐cell RNA sequencing and multiplex immunofluorescence. Increased mesenchymal fibroblasts (FBs) was the main feature in keloid patients. Noticeably, the proportion of pro‐inflammatory FBs was significantly increased in active keloids compared to inactive ones. To explore the nature of proinflammatory FBs, trajectory analysis was conducted and CCN family associated with mechanical stretch exhibited higher expression in active keloids. For vascular endothelial cells (VECs), the proportion of tip and immature cells increased in keloids compared to NS, especially at the periphery of active keloids. Also, keloid VECs highly expressed genes with characteristics of mesenchymal activation compared to NS, especially those from the active keloid center. Multiomics analysis demonstrated the distinct expression profile of active keloids. Clinically, these findings may provide the future appropriate directions for development of treatment modalities of keloids. Prevention of keloids could be possible by the suppression of mesenchymal activation between FBs and VECs and modulation of proinflammatory FBs may be the key to the control of active keloids.
Article
Background Despite the interest in mesenchymal stem cells (MSC), their potential to treat abnormal scarring, especially keloids, is yet to be described. The present study aimed to investigate the therapeutic potential of exosomes derived from human bone marrow MSCs (hBMSC‐Exos) in alleviating keloid formation. Methods Exosomes were isolated from hBMSC, and keloid fibroblasts (KFs) were treated with hBMSC‐Exos. Cell counting kit‐8, wound healing, transwell invasion, immunofluorescence, and western blot assays were conducted to study the malignant phenotype of KFs. Mice were induced with keloids and treated with hBMSC‐Exos. The effect of hBMSC‐Exos on keloid formation in vivo was evaluated by hematoxylin and eosin staining, Masson staining, immunohistochemistry, and western blotting. The GSE182192 dataset was screened for differentially expressed long non‐coding RNA during keloid formation. Next, maternally expressed gene 3 (MEG3) was knocked down in hBMSC to obtain hBMSC‐Exos sh‐MEG3 . The molecular mechanism of MEG3 was investigated by bioinformatic screening, and the relationship between MEG3 and TP53 or MCM5 was verified. Results hBMSC‐Exos inhibited the malignant proliferation, migration, and invasion of KFs at same time as promoting their apoptosis, Moreover, hBMSC‐Exos reduced the expression of fibrosis‐ and collagen‐related proteins in the cells and the formation of keloids caused by KFs. The reduction in MEG3 enrichment in hBMSC‐Exos weakened the inhibitory effect of hBMSC‐Exos on KF activity. hBMSC‐Exos delivered MEG3 to promote MCM5 transcription by TP53 in KFs. Overexpression of MCM5 in KFs reversed the effects of hBMSC‐Exos sh‐MEG3 , leading to reduced KF activity. Conclusions hBMSC‐Exos delivered MEG3 to promote the protein stability of TP53, thereby activating MCM5 and promoting KF activity.
Article
Full-text available
Abnormal regulation of collagen synthesis has been observed in fibroblasts from keloids, benign collagenous tumors that develop as a result of an inherited defect in dermal wound healing. Hydrocortisone reduces the rate of collagen synthesis in fibroblasts from normal adult dermis and scars, but fails to down regulate collagen synthesis in keloid-derived fibroblasts. We show here that loss of glucocorticoid control of collagen synthesis in keloid cells is due to an inability of hydrocortisone to reduce the levels of types I, III, and V collagen mRNA, whereas it coordinately lowers these RNAs in normal adult cells. The defective regulatory mechanism is expressed only in fibroblasts from the lesion. Fibroblasts from uninvolved dermis respond normally to hydrocortisone. Not all glucocorticoid-modulated matrix proteins are abnormally regulated in this disorder; fibronectin mRNA is induced to a similar extent in normal and keloid cells. The failure of hydrocortisone to reduce collagen gene expression is also seen in fibroblasts from fetal dermis. We have reported similarities between keloid and fetal cells with regard to growth factor requirements and growth response to hydrocortisone. Thus, keloids may be due to the inappropriate expression of a pattern of growth and matrix production that is developmentally regulated.
Article
Full-text available
Following the publication of this article, the authors noticed that the published versions of Figs. 2, 7 and 8 contained incorrect western bands. In examining their raw data, the authors realized that they had used the fibroblasts of keloids [high expression of alpha‑smooth muscle actin (α‑SMA)] instead of adult dermal fibroblasts (low expression of α‑SMA) in certain experiments. Note that no significant differences in morphology exist between myofibroblasts (from keloids) and fibroblasts (from normal dermal tissue). These errors were brought to light since the authors identified that the expression of α‑SMA in Fig. 8 was higher compared with that in Fig. 4. After careful scrutiny, they established that the first author, Bin Zhao, who performed the experiments and analyzed the data shown in Figs. 2, 7 and 8, had mislabelled the myofibroblasts as fibroblasts. However, for all the other experiments in the above‑mentioned article, the cells had been used correctly. The authors regret that these errors were featured in the above‑mentioned article, which may possibly have caused confusion for the readers, and the corrected versions of Figs. 2, 7 and 8 are shown opposite and on the next page. These changes did not affect either the results or the conclusions reported in this paper. The authors apologize to the Editor of International Journal of Molecular Medicine and to the readership for any inconvenience caused. [the original article was published in International Journal of Molecular Medicine 39: 153‑159, 2017; DOI: 10.3892/ijmm.2016.2816].
Article
Full-text available
Background: Our understanding of the pathogenesis underlying keloid scar formation is still very limited, and the morphological distinction between hypertrophic and keloid scars remains difficult. Objectives: Using whole biopsy imaging and an objectively quantifiable way to analyze immunoreactivity, we have compared young immature scars (Yscar) with mature normal (Nscar) scars, hypertrophic (Hscar) scars, and keloid (Kscar) scars with their surrounding skin (sNskin). Results: Abnormal scars (Hscar and Kscar) maintain the Yscar phenotype, characterized by a CD34- (tumour biomarker) / α-SMA+ (myofibroblast) dermal region in contrast to Nskin (normal skin), sNskin and mature Nscar which were CD34+/α-SMA-. Yscar, Kscar and Hscar showed abnormal epidermal differentiation (involucrin), but only Hscar and Kscar showed increased epidermal thickness. Yscar did show increased epidermal and dermal proliferation (Ki67), which was absent from abnormal scars where mesenchymal hypercellularity (vimentin) and senescence (p16) were predominant. Keloidal collagen and α-SMA were previously considered to distinguish between Hscar and Kscar. However, α-SMA staining was present in both abnormal scars, while keloidal collagen was mostly present in Kscar. There were no obvious signs of heterogeneity within keloid scars, and sNskin resembled Nskin. Conclusions: Both abnormal scar types (Hscar and Kscar) showed a unique CD34-/α-SMA+/p16+ scar phenotype, but the differences between Hscar and Kscar observed in this study were of a gradient rather than absolute nature. This suggests that scar progression to the mature normal scar phenotype is, for as yet unknown reasons, hindered in hypertrophic and keloid scars. This article is protected by copyright. All rights reserved.
Article
Full-text available
Keloid disorder (KD) is a fibroproliferative condition caused by dysregulated wound healing following wounding of the skin. The pathogenesis of KD has not been fully elucidated and current treatment is unsatisfactory. There is increasing evidence of the role of stem cells in KD. This review discusses the role of embryonic stem (ESC)-like cells and mesenchymal stem cells in the pathogenesis of KD. It is proposed that dysfunction of the ESC-like population localized to the endothelium of the microvessels and perivascular cells within the keloid-associated lymphoid tissues may give rise to the aberrant fibroblasts and myofibroblasts via a mesenchymal stem cell intermediate in keloid lesions, by undergoing an endothelial-to-mesenchymal transition. We also discuss the role of the renin-angiotensin system (RAS), the immune system, and the inflammatory response, on stem cell proliferation and differentiation. The understanding of the precise roles of these stem cells and interplay of the associated regulatory pathways could lead to the development of targeted therapy for this enigmatic and challenging condition. The demonstration of the expression of components of the RAS and cathepsins B, D, and G that constitute bypass loops of the RAS, by the ESC-like population, suggests that the primitive population may be a therapeutic target by modulation of the RAS, using existing medications.
Article
Full-text available
Keloid scars are described as benign fibro‐proliferative dermal outgrowths that commonly occur in pigmented skin post cutaneous injury, and continue to grow beyond the boundary of the original wound margin. There is a lack of thorough understanding of keloid pathogenesis and thus keloid therapeutic options remain ill‐defined. In view of the poor response to current therapy and high recurrence rates, there is an unmet need in improving our knowledge and therefore in identifying targeted and effective treatment strategies in management of keloids. Keloid research however, is hampered by a lack of relevant animal models as keloids do not spontaneously occur in animals and are unique to human skin. Therefore, developing novel animal models and non‐animal models for functional evaluation of keloid cells and tissue for better understanding their pathobiology and response to putative candidate therapies are essential. Here, we present the key concepts and relevant emerging research on two‐dimensional (2D), and 3D cell and tissue models for functional testing of keloid scars. We will describe in detail current models including in‐vitro mono‐ and co‐cultures, multi‐cellular spheroids (organoids) and organotyopic cultures, ex‐vivo whole skin keloid tissue organ culture models as well as in‐vivo human patient models. Finally, we discuss role played by time as 4D in sequential temporal approach in the latter model. The use of these unique models will no doubt prove pivotal in identification of new drug targets as well as biomarkers, in testing for emerging novel therapeutics, and in enhancing our understanding of keloid disease biology. This article is protected by copyright. All rights reserved.
Article
There is no good evidence on clinical grounds for any definite differentiation between keloids and hypertrophic scars, employing any of the numerous parameters. Incidence and theories of keloid formation are presented at length with a chemo-therapeutic regimen.
Article
Background: Fibroproliferative disorders result in excessive scar formation, are associated with high morbidity, and cost billions of dollars every year. Of these, keloid disease presents a particularly challenging clinical problem because the cutaneous scars progress beyond the original site of injury. Altered mechanotransduction has been implicated in keloid development, but the mechanisms governing scar progression into the surrounding tissue remain unknown. The role of mechanotransduction in keloids is further complicated by the differential mechanical properties of keloids and the surrounding skin. Methods: The authors used human mechanical testing, finite element modeling, and immunohistologic analyses of human specimens to clarify the complex interplay of mechanical stress, strain, and stiffness in keloid scar progression. Results: Changes in human position (i.e., standing, sitting, and supine) are correlated to dynamic changes in local stress/strain distribution, particularly in regions with a predilection for keloids. Keloids are composed of stiff tissue, which displays a fibrotic phenotype with relatively low proliferation. In contrast, the soft skin surrounding keloids is exposed to high mechanical strain that correlates with increased expression of the caveolin-1/rho signaling via rho kinase mechanotransduction pathway and elevated inflammation and proliferation, which may lead to keloid progression. Conclusions: The authors conclude that changes in human position are strongly correlated with mechanical loading of the predilection sites, which leads to increased mechanical strain in the peripheral tissue surrounding keloids. Furthermore, increased mechanical strain in the peripheral tissue, which is the site of keloid progression, was correlated with aberrant expression of caveolin-1/ROCK signaling pathway. These findings suggest a novel mechanism for keloid progression.
Article
Excess scar formation can occur after skin injurふy and lead to abnormal scar formation, such as keloids and hypertrophic scars, which are characterised by substantial deposition of extracellular matrix in the dermis. Periostin, an extracellular matrix protein that plays a crucial role in skin development and maintaining homeostasis, is also involved in skin disorders such as systemic/limited scleroderma, wound closure, and abnormal scar formation. However, the mechanism of periostin involvement in abnormal scar formation is not yet fully understood. In this study, we investigated the mechanism by which periostin is involved in abnormal scar formation. Treatment of human dermal fibroblasts (HDFs) with IL-4 and IL-13, which are cytokines of Th2 type immune responses that are up-regulated in abnormal scars, dramatically elevated the levels of periostin mRNA and protein, and also promoted the secretion of periostin by HDFs. Transforming growth factor-β1 (TGF-β1) had the same effect on HDFs as IL-4 and IL-13. Stimulation of HDFs with periostin promoted RhoA/ROCK pathway-mediated TGF-β1 secretion from HDFs. Our results suggest that IL-4 and IL-13 induce periostin expression and secretion, and in turn, secreted periostin induces RhoA/ROCK pathway-mediated TGF-β1 secretion. Secreted TGF-β1 then induces further periostin production and secretion, thereby promoting abnormal scar formation.