You are on page 1of 11

RESEARCH ARTICLE

VITILIGO

CXCL10 Is Critical for the Progression and Maintenance


of Depigmentation in a Mouse Model of Vitiligo
Mehdi Rashighi,1 Priti Agarwal,1 Jillian M. Richmond,1 Tajie H. Harris,2* Karen Dresser,3
Ming-Wan Su,4 Youwen Zhou,4 April Deng,3 Christopher A. Hunter,2 Andrew D. Luster,5 John E. Harris1†
Vitiligo is an autoimmune disease of the skin that results in disfiguring white spots. There are no U.S. Food and Drug
Administration–approved treatments for vitiligo, and most off-label treatments yield unsatisfactory results. Vitiligo
patients have increased numbers of autoreactive, melanocyte-specific CD8+ T cells in the skin and blood, which are
directly responsible for melanocyte destruction. We report that gene expression in lesional skin from vitiligo
patients revealed an interferon-g (IFN-g)–specific signature, including the chemokine CXCL10. CXCL10 was elevated
in both vitiligo patient skin and serum, and CXCR3, its receptor, was expressed on pathogenic T cells. To address the
function of CXCL10 in vitiligo, we used a mouse model of disease that also exhibited an IFN-g–specific gene sig-
nature, expression of CXCL10 in the skin, and up-regulation of CXCR3 on antigen-specific T cells. Mice that received
Cxcr3−/− T cells developed minimal depigmentation, as did mice lacking Cxcl10 or treated with CXCL10-neutralizing

Downloaded from stm.sciencemag.org on May 3, 2015


antibody. CXCL9 promoted autoreactive T cell global recruitment to the skin but not effector function, whereas
CXCL10 was required for effector function and localization within the skin. Surprisingly, CXCL10 neutralization in
mice with established, widespread depigmentation induces reversal of disease, evidenced by repigmentation.
These data identify a critical role for CXCL10 in both the progression and maintenance of vitiligo and thereby
support inhibiting CXCL10 as a targeted treatment strategy.

INTRODUCTION Topical steroids are impractical when large surface areas are affected,
Vitiligo is a disease of the skin that afflicts ~0.5 to 2% of the population and they carry the risk of systemic absorption with suppression of the
and results in prominent, disfiguring white spots that may become adrenal axis, in addition to striae and skin atrophy (1). Topical calci-
widespread (1). Vitiligo pathogenesis incorporates both intrinsic defects neurin inhibitors appear to have fewer risks, but they are costly and less
within melanocytes that activate the cellular stress response and auto- effective and treat only focal disease (17). Systemic immunosuppression
immune mechanisms that target these cells (2–12). Patients with vitiligo has been reported to halt the progression of vitiligo and reverse disease
have increased numbers of autoreactive, melanocyte-specific CD8+ T through repigmentation (18); however, the risks associated with a non-
cells in the skin and blood (13, 14). T cells infiltrate the skin during targeted systemic approach are substantial and usually deemed un-
vitiligo and localize to the epidermis where melanocytes, their target acceptable when compared to the benefits. Therefore, the identification
cells, reside. In lesional skin, CD8+ T cells are found in close proximity of targeted therapies with an improved safety profile would be a substan-
to dying melanocytes (15, 16), and one study reported that melanocyte- tial advance. Recent progress in understanding the key cytokines that
specific, CD8+ T cells isolated from lesional skin migrated into non- promote psoriasis and related autoimmune diseases (such as rheumatoid
lesional skin ex vivo, found their melanocyte targets in situ, and arthritis, inflammatory bowel disease, and others) have resulted in
killed them (13). The melanocyte niche in the basal layer of the ep- biological treatments with excellent efficacy and safety profiles, substan-
idermis is not vascularized, so T cells that migrate into the skin in tially improving patients’ quality of life (19). The cytokines that drive
vitiligo must efficiently navigate through the dermis to find their targets vitiligo pathogenesis have not been well defined, although they are
and mediate depigmentation. The signals that promote melanocyte- not likely shared with these diseases because biological treatments
specific T cell migration into and through the skin during vitiligo for psoriasis have been ineffective for vitiligo (20–22).
are unknown. The dynamics at play within established vitiligo lesions are unknown,
There are currently no U.S. Food and Drug Administration–approved although depigmented skin is widely thought to be inactive, such that
medical treatments for vitiligo, and available off-label treatments are T cells have left the tissue and melanocytes are unable to regenerate.
problematic and often ineffective. Narrow-band ultraviolet B (nbUVB) However, the ability of even long-standing lesions to repigment after
light is only provided in select facilities and requires time away from treatment suggests that this is not always the case, but rather an ac-
work and school multiple times per week, resulting in lost productivity. tive immune process is at work to maintain depigmentation. Further
support for this hypothesis is found in the rapid repigmentation
1
Division of Dermatology, Department of Medicine, University of Massachusetts of lesional human vitiligo skin after transplantation onto nude mice
Medical School, Worcester, MA 01605, USA. 2Department of Pathobiology, University (23), suggesting that host factors play an important role in maintain-
of Pennsylvania School of Veterinary Medicine, Philadelphia, PA 19104, USA. ing depigmentation.
3
Department of Pathology, University of Massachusetts Medical School, Worcester,
MA 01605, USA. 4Department of Dermatology and Skin Science, University of British
Here, we report that both human vitiligo patients and a mouse model
Columbia, Vancouver, British Columbia V5Z 4E8, Canada. 5Center for Immunology and of vitiligo reflect an interferon-g (IFN-g)–specific T helper 1 (TH1)
Inflammatory Diseases, Division of Rheumatology, Allergy and Immunology, Massa- cytokine signature in the skin that includes the IFN-g–dependent chemo-
chusetts General Hospital, Harvard Medical School, Boston, MA 02114, USA. kines CXCL9, CXCL10, and CXCL11. CXCL10 is highly expressed
*Present address: Department of Neuroscience, School of Medicine, University of
Virginia, Charlottesville, VA 22908, USA. in the skin and serum of patients with vitiligo, and it promotes auto-
†Corresponding author. E-mail: john.harris@umassmed.edu reactive T cell localization to the epidermis to initiate vitiligo in our

www.ScienceTranslationalMedicine.org 12 February 2014 Vol 6 Issue 223 223ra23 1


RESEARCH ARTICLE

mouse model. Targeting CXCL10 therapeutically in mice not only A


prevents disease development but also induces repigmentation in 64
established lesions, implicating this chemokine as a critical player
in both the progression and maintenance of vitiligo. Our results sup-
16

Relative fold expression


port further exploration of CXCL10 neutralization as a targeted sys-
temic therapeutic approach in the treatment of vitiligo.
4

D RP A

1
TYCT 1
TY AN

PM
TRR
RESULTS

L
M
1
Gene expression in vitiligo reflects a TH1-mediated
immune response
–4

C L2

C L
C CL10
IN CL11
U DO 9
MBD
M 1
IFX2
C
C L1
C L2

C L5
C CR

IFIT1
IC 2
IC M
VC M
To identify key cytokine pathways in vitiligo, we measured the gene

XC 3
C 2
C 0
X

X
X
C
C 7

IT
A
A 1
A2
M
expression profile in lesional skin of vitiligo patients. Our criteria for

1
selecting skin specimens relied on the presence of T cell infiltrates to TH2
–16 TH17
determine inflammatory gene signatures in active lesions. Lesional TH1 Type II
biopsies frequently miss the T cells because visible depigmentation of IFN Type I Adhesion
IFN
the skin does not become apparent until long after melanocyte death, –64

Downloaded from stm.sciencemag.org on May 3, 2015


up to 48 days in humans (24). Consequently, we screened a clinical
B C D
archive of formalin-fixed, paraffin-embedded (FFPE) biopsies for those
CXCL9 CXCL10 CXCL11
that contained a T cell infiltrate, selected samples from five vitiligo pa- ns
tients and five age- and site-matched controls, and analyzed their RNA 4000
3000
4000
3000
* 4000
3000 ns
expression profile using the Illumina DASL assay. We observed signif- 2000
2000 2000
icant loss of melanocyte-specific transcripts in lesional skin compared

pg/ml
pg/ml

pg/ml
1000 1000 1000
to healthy controls, validating this approach. Chemokine expression 300 300 300
revealed a predominantly TH1-mediated immune response (Fig. 1A). 200 200 200
On the basis of a small panel of genes whose expression is more depen- 100 100 100
dent on either type I IFN (IFN-a/b) or type II IFN (IFN-g), we identified 0 0 0

y
a distinctly IFNG-specific signature (Fig. 1A).

o
y

th
y

ig
lth

ig

lth

ig

til
ea
IFN-g has previously been shown to be induced in human vitiligo
til

til
ea

ea

Vi
Vi

H
Vi
H

H
(13, 25), and it is required for the development of depigmentation in a
mouse model of vitiligo (26). IFN-g signaling induces the expression Fig. 1. Vitiligo patients express IFN-g–dependent chemokines in
of more than 200 different gene targets (27); however, our earlier re- skin and blood. (A) We analyzed the gene expression profiles within
sults suggested that IFN-g was specifically required for T cell accumu- skin biopsies from vitiligo patients compared to age- and site-matched
lation within the skin (26). Therefore, IFN-g–dependent chemokines, controls using quantile normalization (n = 5 per group). Melanocyte-
specific transcripts were significantly reduced, consistent with melano-
which induce T cell homing into peripheral tissues, and endothelial
cyte destruction in vitiligo. Chemokine expression reflected a T H1
adhesion molecules, which promote transmigration into tissues, were profile, and the expression of specific gene targets of either type I or
top candidates for this role (28, 29). We found that the chemokines type II IFNs revealed an IFN-g–dominant response. The expression of
CXCL9, CXCL10, CXCL11, and CCL5 were highly induced in vitiligo, genes capable of T cell recruitment to the skin reflected expression of
whereas the adhesion molecules ICAM1, ICAM2, and VCAM1 were chemokines, rather than adhesion molecules. (B to D) Serum samples
not (Fig. 1A). We confirmed these expression data in an additional 11 from vitiligo donors or healthy controls were analyzed by enzyme-
samples (8 vitiligo samples, 3 controls) using NanoString nCounter linked immunosorbent assay (ELISA) to determine levels of CXCL9 (B),
profiling (fig. S1A). CXCL10 (C), and CXCL11 (D). CXCL10 was significantly elevated, whereas
Previous studies reported elevated levels of CXCL9 and CXCL10 CXCL9 and CXCL11 were not (n = 16 to 32 donors assayed in duplicate;
in the serum of patients with autoimmune thyroiditis and primary *P = 0.0148, two-tailed Mann-Whitney U test). ns, not significant.
adrenal insufficiency (30, 31). To determine whether these chemo-
kines could be detected in vitiligo patients as well, we measured serum
from vitiligo patients and healthy controls using ELISA. We found vitiligo patients and five HLA-A2+ healthy controls using melanocyte
that CXCL10 was indeed significantly elevated in vitiligo patients, antigen–specific HLA pentamers (gp100 and tyrosinase) to identify
but CXCL9 and CXCL11 were not significantly different from healthy autoreactive CD8+ T cells (Fig. 2A). Consistent with previous studies
controls (Fig. 1, B to D). (13, 14), ~0.5 to 1% of CD8+ T cells were positive for each pentamer in
vitiligo patients, a significant difference from healthy controls (fig. S2).
CXCR3 is expressed on autoreactive T cells We found that most of the pentamer+ cells in vitiligo patients ex-
To examine the potential for autoreactive T cells to respond to IFN-g– pressed CXCR3, unlike in healthy controls (Fig. 2B). To determine
associated chemokines expressed in the skin of patients with vitiligo, whether CXCR3 is expressed in the lesional skin of patients with
we measured the expression of CXCR3, the common receptor for vitiligo, we used immunohistochemistry on skin biopsies from four
CXCL9, CXCL10, and CXCL11, on melanocyte-specific CD8+ T cells. patients with vitiligo, which revealed CXCR3 expression within
We analyzed the blood of five human leukocyte antigen (HLA)–A2+ each sample (Fig. 2C and fig. S3).

www.ScienceTranslationalMedicine.org 12 February 2014 Vol 6 Issue 223 223ra23 2


RESEARCH ARTICLE

gp100 Tyr None and Cxcl10. Both chemokines were ex-


A
0.40% 0.46% 0%
B
6 ** pressed in mouse skin after vitiligo in-

Ag-specific CD8+ T cells


Ratio CXCR3+/CXCR3–
duction and were dependent on IFN-g
for their expression (Fig. 3A and fig. S1,
Vitiligo B and C). Flow cytometric analysis of cells
4
from the blood, spleen, and skin-draining
0.11% 0.15% 0% lymph nodes of mice with vitiligo revealed
0.01% that a substantial proportion of autoreac-
CXCR3

0.07% 0.11%
2 tive T cells express CXCR3 (Fig. 3B, fig. S4,
and table S1), similar to humans with dis-
Healthy ease. Thus, expression of CXCL9, CXCL10,
and CXCR3 strongly associated with the
0.07% 0.06% 0.02% 0 depigmentation in vitiligo in both humans
Healthy Vitiligo and mice, and so we tested their functional
Pentamer
significance in our mouse model.
C Subject 1 Subject 2 Subject 3 Subject 4
CXCR3 expression on T cells
is required for the development

Downloaded from stm.sciencemag.org on May 3, 2015


of vitiligo
CD3 To test whether CXCR3 is required for
the development of depigmentation, we
adoptively transferred either wild-type or
Cxcr3−/− PMEL T cells to induce vitiligo.
CD8 Cxcr3−/− T cells were impaired in their abil-
ity to induce depigmentation (Fig. 4, A
and B) and failed to accumulate in the skin
(Fig. 4C), despite normal numbers within
the skin-draining lymph nodes (Fig. 4D).
CXCR3 This pattern is similar to that after treat-
ment with IFN-g–neutralizing antibody
(26) and supports the observation that
Cxcr3−/− T cell receptor (TCR) transgenic
Fig. 2. CXCR3 is expressed on antigen-specific cells in the blood and CD8+ T cells in the skin
host mice are protected from vitiligo in a
lesions of vitiligo patients. (A) Expression of CXCR3 on melanocyte antigen–specific pentamer+ CD8+
+ +
T cells in the blood of HLA-A2 patients with vitiligo, pregated on CD8 T cells. Flow plots representative of
separate mouse model of hair depigmenta-
five vitiligo patients and five healthy controls are shown; gp100-specific pentamer (gp100), tyrosinase- tion (33). To quantify −/−
the decreased effi-
+ − ciency of Cxcr3 T cell migration to the
specific pentamer (Tyr), and control, no pentamer (None). (B) The ratio of CXCR3 /CXCR3 antigen-specific
T cells is significantly higher in vitiligo patients than in healthy controls (n = 5 donors with two pentamers skin within the same host, we adoptively
each; **P = 0.0098, two-tailed unpaired t test). (C) Immunohistochemical staining of CD3, CD8, and CXCR3 transferred equal numbers of cyan fluores-
in the skin of four patients with active vitiligo, identified from an FFPE clinical tissue bank. All showed cent protein–positive (CFP+) wild-type PMEL
positive staining for CD3, CD8, and CXCR3. Scale bars, 100 mm. T cells and green fluorescent protein–positive
(GFP+) wild-type or Cxcr3−/− PMEL T cells
into hosts and measured their total numbers
A mouse model of vitiligo reflects the IFN-g signature in the skin and skin-draining lymph nodes 5 weeks after transfer using
observed in patient samples flow cytometry. On average, wild-type T cells were three times more
To interrogate the role of CXCR3 chemokines in vitiligo, we used a prevalent than Cxcr3−/− T cells in the skin, despite similar numbers in
mouse model that develops depigmentation of the epidermis, similar skin-draining lymph nodes (Fig. 4E and fig. S5).
to human disease (26). Unlike wild-type mice, Krt14-Kitl* mice have
increased numbers of epidermal melanocytes, similar to human skin CXCL9, CXCL10, and CXCL11 do not act
(32). After adoptive transfer of premelanosome protein-specific CD8+ redundantly in vitiligo
T cells (PMELs) and in vivo activation with recombinant vaccinia virus Because CXCL9, CXCL10, and CXCL11 all signal through the CXCR3
expressing their cognate antigen (PMEL), Krt14-Kitl* host mice develop receptor, we sought to determine whether individual chemokines were
patchy epidermal depigmentation on their ears, tails, noses, and foot- required for depigmentation, or whether they act redundantly, by
pads (26). Gene expression profiling of lesional skin from mice with testing chemokine-deficient hosts in our model. Both Cxcl9−/− and
vitiligo revealed a similar chemokine signature to human vitiligo, Cxcl10−/− mice were originally created on the 129 mouse strain and
with a minimal type I IFN gene signature (Fig. 3A). Cxcl11, often co- then subsequently backcrossed to C57BL/6. Because of close genetic
expressed with Cxcl9 and Cxcl10, is not expressed in the C57BL/6 linkage of Cxcl9, Cxcl10, and Cxcl11, this approach results in Cxcl9−/−
mouse strain because of a spontaneous null mutation (28), and so fur- and Cxcl10−/− C57BL/6 mice with an intact Cxcl11 gene, unlike the
ther studies in our mouse model focused on the contributions of Cxcl9 parent strain (28). Therefore, when either knockout mouse is tested

www.ScienceTranslationalMedicine.org 12 February 2014 Vol 6 Issue 223 223ra23 3


RESEARCH ARTICLE

A ing chemokine (34). However, despite the decreased recruitment effi-


64 ciency of PMELs in Cxcl9−/− hosts, the PMELs were still highly
functional because they were capable of inducing depigmentation
comparable to wild-type hosts. In contrast, Cxcl10−/− hosts had simi-
Relative fold expression

16
lar numbers of PMELs in the dermis compared to wild-type hosts
(Fig. 5C) and a trend toward a lower number of cells in the epider-
mis (Fig. 5D), suggesting that CXCL10 is required for T cell function
within the skin beyond simple recruitment and may play a role in
4 directed migration within the skin.
CXCL10 has also been reported to play an important role in T cell
A

1
C 1
TY AN

PM

tethering and activation (34). We therefore assessed whether the cells


D RP
T
TRR
L

TY
M

1 responded to antigen and target cells in the same capacity in our wild-
type and Cxcl10−/− mice. Using CD69 as a marker of skin memory T
IN CL 0
U O
M D
M 1
IF 2
IF T1

VC M2
C L2
C L2
C L5
C CR
C L

IC 2
IC M1
C
C L1

X 1
C 2
C 0
X
XC 3

B
C
C 7

X
X
I
IT

A
D 9

A
cells that had been previously activated in vivo (35), we addressed

A
M
1
TH2 whether or not epidermal PMELs maintained high CD44 expression
–4 TH17 TH1
Type II Type I (Fig. 5, E and F). We found that the ratio of CD44lo to CD44hi epi-
IFN Adhesion dermal CD69+ PMELs was significantly higher in Cxcl10−/− hosts, sug-
IFN
gesting that they have an altered activation status compared to PMELs

Downloaded from stm.sciencemag.org on May 3, 2015


B Blood Spleen Lymph node in wild-type mice (Fig. 5E). This ratio was not elevated in the lymph
nodes of host mice, indicating a skin-specific effect (Fig. 5F).
7.7% 5.6% 5.1%
CXCR3

Neutralization of CXCL10 both prevents and reverses


depigmentation in vitiligo
Cxcl9−/− and Cxcl10−/− hosts have previously been shown to exhibit
defects in T cell priming against viruses (34, 36, 37). Therefore, to
10.8% 3.8% 1.2%
avoid defective priming of PMELs in our model and to assess the roles
GFP of these chemokines during the effector phase of disease in wild-type
hosts, we treated mice with either CXCL9- or CXCL10-neutralizing
Fig. 3. Expression of CXCL9, CXCL10, and CXCR3 in a mouse model
antibody or phosphate-buffered saline (PBS) control beginning 2 weeks
of vitiligo. (A) Gene expression in fresh ear skin from mice 5 weeks after
vitiligo induction and controls were analyzed via microarray using quan-
after adoptive transfer. This time frame thereby allowed for normal
tile normalization (n = 3 per group). As in human disease, mice activation of the PMEL T cells and clearance of the virus before chemo-
displayed a TH1-specific response in the skin and expressed the IFN- kine neutralization. Consistent with our observations in chemokine-
g–dependent chemokines CXCL9, CXCL10, and CCL5. (B) CXCR3 was deficient hosts, neutralization of CXCL10 significantly reduced
expressed on melanocyte-specific CD8+ T cells (GFP+ PMEL) in the depigmentation in our model, whereas neutralization of CXCL9 did
blood, spleen, and skin-draining lymph nodes from mice with vitiligo, not (Fig. 6A).
gated on CD8+ T cells (representative flow plots, n = 4 mice from three To test the efficacy of CXCL10 neutralization as a treatment of es-
experiments). tablished disease, vitiligo was induced as above. Eight to 10 weeks later,
about 10% of mice exhibited >50% depigmentation of the tails, and we
randomized them into two groups: one treated with CXCL10-neutralizing
on the C57BL/6 background, Cxcl11 expression in the knockout strain antibody and a control group treated with PBS (Fig. 6B) or isotype con-
could potentially compensate for the deficient chemokine. In our trol antibody (fig. S6). About 4 weeks after initiation of treatment, mice
mouse model of vitiligo, Cxcl10−/− hosts developed minimal de- receiving CXCL10-neutralizing antibody began to develop repigmenta-
pigmentation (Fig. 5, A and B) similar to when Cxcr3−/− T cells were tion, which continued for an additional 4 weeks, whereas control mice
tested, suggesting that vitiligo depends on this single chemokine and primarily exhibited minimal improvement or progression. Repigmen-
that CXCL11 cannot substitute for its function. In contrast, Cxcl9−/− tation occurred from the hair follicles on the tail (Fig. 6C).
hosts developed depigmentation comparable to wild-type controls.

CXCL10 promotes epidermal positioning and effector DISCUSSION


function of autoreactive T cells
To assess the functional contributions of CXCL10 to depigmentation, The recent success of systemic biological therapies that target cytokines
we analyzed the positioning of PMEL T cells in the layers of the skin in psoriasis and other inflammatory diseases suggests that a similar
and their activation status 5 to 6 weeks after transfer. Because melano- approach might be effective for vitiligo. Our observation that the im-
cytes are located in the epidermis, we first addressed whether or not mune response in vitiligo is TH1-specific is consistent with a previous
PMELs could migrate into the epidermis in chemokine knockout mice report (38) and may explain why targeted treatments for psoriasis, a
using flow cytometry to separately analyze the dermal and epidermal prototypic TH17-mediated disease, are ineffective for vitiligo (20–22).
layers of the skin. Cxcl9−/− hosts exhibited lower numbers of PMELs The identification of an IFNG-specific signature in vitiligo is in sharp con-
in both the dermis (Fig. 5C) and epidermis (Fig. 5D) compared to trast to other inflammatory skin diseases, including dermatomyositis,
wild-type hosts, supporting its role as a global positioning and recruit- cutaneous lupus erythematosus, and lichen planus, which express an

www.ScienceTranslationalMedicine.org 12 February 2014 Vol 6 Issue 223 223ra23 4


RESEARCH ARTICLE

A WT PMEL transfer depigmentation, implicates this chemo-


kine pathway as being functionally required
for vitiligo. Therefore, small-molecule in-
hibitors of CXCR3, currently in develop-
CXCR3–/– PMEL transfer ment by multiple sources (41), may be
effective therapies. The clear dominance
of CXCL10 over CXCL9 in our mouse
model of vitiligo is intriguing, providing
an opportunity to target CXCL10 alone
B C D for developing new treatments. Depend-
500
#PMEL/104 CD45+ cells
*** 250 ns ing on the model of inflammation studied,

#PMEL/104 CD45+ cells


15
*** 400 200
the CXCR3 ligands can have redundant,
dominant, cooperative, or even antago-
Vitiligo score

10 300 150 nistic functions (28). CXCL10 dominance


has been reported primarily during viral
200 100 infections of peripheral tissues (28), but al-
5 so in the interleukin (IL)–10null model of
100 50
inflammatory bowel disease (42). In the

Downloaded from stm.sciencemag.org on May 3, 2015


0 0 0 MRL-lpr lupus mouse model, Cxcl9−/−
–/– –/–
Control WT CXCR3 WT CXCR3–/– WT CXCR3 mice were protected from disease, whereas
Cxcl10−/− mice were not, suggesting CXCL9
E Fig. 4. CXCR3 expression on T cells is re- dominance in that system (43). Other mod-
6
* quired for the development of vitiligo
in mice. (A) Vitiligo was induced through
els, including brain inflammation after
Frequency ratio

malarial infection (44) and herpes sim-


adoptive transfer of wild-type (WT) or
4 Cxcr3 −/−
PMELs. Hosts receiving Cxcr3 −/− plex virus-2 infection of the spinal cord (45),
PMELs developed less depigmentation required both CXCL9 and CXCL10 for
than those that received WT PMELs (rep- T cell recruitment, suggesting that spatio-
2 ns resentative images of tail depigmentation temporal differences in chemokine expres-
shown). (B) The degree of depigmenta- sion may be responsible for a cooperative,
tion was scored blindly by one observer coordinated role of these chemokines (28).
5 weeks after vitiligo induction. Disease The fact that mice on the C57BL/6 back-
0 scores were significantly lower in mice that
WT/WT WT/CXCR3 –/– WT/WT WT/CXCR3 –/– ground do not express CXCL11 suggests
received Cxcr3−/− PMELs compared to
that vitiligo in our mouse model does not
Skin Lymph nodes those that received WT PMELs [one-way
depend on this chemokine. The presence
analysis of variance (ANOVA) with Tukey’s −/−
post-tests: control versus WT, P < 0.0001; WT versus Cxcr3−/−, P < 0.0001]. (C) There was impaired ac- of intact Cxcl11 in Cxcl9 mice did not −/−
cumulation of Cxcr3−/− PMELs in the ear skin 5 to 6 weeks after vitiligo induction (P < 0.0001, two-tailed exacerbate the disease, and in Cxcl10
unpaired t test). (D) However, there was no significant difference in the skin-draining lymph nodes. (A to mice, Cxcl11 was unable to compensate for
D) Data were pooled from three independent experiments (n = 30 mice per group). (E) WT CFP+ PMELs its absence. These data suggest that CXCL11
and WT or Cxcr3−/− GFP+ PMELs were cotransferred in equal proportions into hosts to induce vitiligo. does not play a major role in our mouse
The ratio of WT/WT or WT/Cxcr3−/− PMELs (frequency ratio) in the skin and skin-draining lymph nodes model of vitiligo; however, we cannot yet rule
was analyzed by flow cytometry. The frequency ratio for WT/Cxcr3−/− was significantly higher than that out a role in human disease. Additional
for WT/WT in the skin, but equal frequency ratios were found in skin-draining lymph nodes, indicating studies will be required to determine which
impaired skin homing of Cxcr3−/− PMELs (n = 4 to 5 mice per group, three experiments; representative
cell types produce CXCL9 and CXCL10 in
experiment shown; skin: P = 0.0303, two-tailed unpaired t test).
vitiligo, how they interact to promote vitiligo
pathogenesis, and the relative importance
of each chemokine in humans.
IFN signature that reflects both type I (IFNA/B) and type II (IFNG) Our observations suggest that CXCL10 is required for more than T
IFNs, with a predominance of type I over type II (39, 40). Consistent cell recruitment to the target tissue. This is consistent with recently
with our findings, other studies reported increased expression of IFNG described functional roles for CXCL10 within infected tissues and
in vitiligo skin; however, they did not measure the expression of down- lymph nodes beyond simple recruitment, including proper localiza-
stream gene targets or their functional/biological significance (13, 25, 38). tion within the lymph node microenvironment (34, 36), increased mi-
We previously reported that neutralizing IFN-g in a mouse model of gration velocity in the brain (46), promoting T cell interaction with
vitiligo prevented depigmentation (26). Together, these data suggest that dendritic cells (34), and lymph node retention (47). Any or all of these
targeting IFN-g–dependent events may be a particularly effective treat- CXCL10-dependent functions may also be important in vitiligo. The de-
ment strategy for vitiligo. creased localization of autoreactive T cells to the epidermis in Cxcl10−/−
The expression of CXCR3 on autoreactive T cells both in human mice with relatively normal numbers in the dermis suggests that
vitiligo patients and in our mouse model, as well as the requirement of proper localization within the skin is one of these important func-
CXCR3 for autoreactive T cell accumulation in the skin and subsequent tions, in contrast to CXCL9 deficiency, which reveals a more global

www.ScienceTranslationalMedicine.org 12 February 2014 Vol 6 Issue 223 223ra23 5


RESEARCH ARTICLE

** A B

Fold change over baseline


A B *
15 *** ns
ns
**
15 1.5
Control

Vitiligo score

Vitiligo score
WT 10 1.0
10
–/–
Cxcl9
5 5 0.5
Cxcl10 –/–
0 0 0.0
PBS CXCL10 Ab

Tx

b
–/

–/

ol
l

A
ro

tr
9

10

L9

0
nt

CL

on

L1
CL

XC
Co

C
CX

XC
CX

C
C D C
Dermis Epidermis
Before treatment
# PMEL per 104 CD45+ cells

# PMEL per 104 CD45+ cells

ns ns
1000 * 8000 *
800 6000 After treatment

Downloaded from stm.sciencemag.org on May 3, 2015


600
4000
400
200 2000 Fig. 6. CXCL10-neutralizing antibody both prevents and reverses vitiligo.
0
(A) Vitiligo was induced, and mice either received no treatment or were treated
0
with PBS (No Tx, n = 21), CXCL9-neutralizing antibody (CXCL9 Ab, n = 15), or


0 –/–

0 –/–
–/

–/
T

CXCL10-neutralizing antibody (CXCL10 Ab, n = 15). Antibodies were


L9

L9
W

W
L1

1
CL
XC

C
XC

administered three times weekly beginning 2 weeks after vitiligo induction.


CX

CX
C

E F The degree of depigmentation was scored blindly by one observer 5 weeks


Epidermis Lymph nodes
after vitiligo induction (after 3 weeks of treatment). CXCL10 antibody–treated
Ratio CD44lo/CD44hi

Ratio CD44lo/CD44hi

2.5 ns
0.6 * mice had significantly lower disease scores than control mice (pooled from
ns 2.0
CD69+ PMEL
CD69+ PMEL

three independent experiments; one-way ANOVA P = 0.0012 with Tukey’s


0.4 1.5 post-tests: control versus no treatment, P = 0.0011; no treatment versus
1.0 CXCL10 antibody, P = 0.0273). (B and C) Mice with extensive vitiligo
0.2 (>50% depigmentation on the tail) were treated with CXCL10-neutralizing
0.5 antibody (CXCL10 Ab, n = 7) or vehicle control (PBS, n = 3) for 8 weeks. (B)
0.0 0.0 Degree of repigmentation in the tails. Photographs of mouse tails before and


0 –/–

0 –/–

8 weeks after treatment were analyzed using ImageJ software to quantify


–/

–/
T

T
L9

L9
W

W
1

the extent of repigmentation. Fold change over baseline reflects the total
CL

CL
XC

XC
CX

CX
C

pigmentation after treatment divided by total pigmentation before treat-


ment; therefore, a score of 1.0 reflects no change. CXCL10 antibody treatment
Fig. 5. CXCL9 and CXCL10 play nonredundant roles in vitiligo: CXCL9
resulted in repigmentation of the tails, whereas control mice experienced
induces global recruitment, whereas CXCL10 promotes epidermal
disease progression (one-tailed unpaired t test, **P = 0.0019). (C) Repigmen-
positioning and activation status. Vitiligo was induced in WT, Cxcl9−/−,
tation in a mouse with vitiligo after 8 weeks of treatment with CXCL10 anti-
or Cxcl10−/− hosts. (A) Representative pictures of tail depigmentation. (B)
body (representative example shown).
The degree of depigmentation was scored blindly 5 to 6 weeks after vitiligo
induction. Cxcl10−/− mice had significantly lower disease scores than WT
mice (n = 27 control, 80 WT, 44 Cxcl9−/−, 59 Cxcl10−/− mice from nine defect in skin homing despite normal effector function. The appearance
separate experiments; one-way ANOVA P < 0.0001 with Tukey’s post-tests:
of CD69+CD44lo autoreactive T cells within the skin of Cxcl10−/− hosts
control versus WT, P = 0.0003; control versus Cxcl9−/−, P = 0.0002; WT versus
Cxcl10−/−, P = 0.0009; Cxcl9−/− versus Cxcl10−/−, P = 0.0007). PMEL accumu- suggests that CXCL10 may be required for maintained expression of this
lation in the skin was analyzed by flow cytometry of the dermis and epi- memory marker within the skin. Alternatively, CXCL10 may antago-
dermis. (C) Reduced number of PMELs in the dermis of Cxcl9−/−, but not nize the development of a CD44lo population of CD8+ T cells, which was
Cxcl10−/−, mice (one-way ANOVA P = 0.0219 with Tukey’s post-tests: WT shown to be capable of suppressing autoimmune responses in the gut (48).
versus Cxcl9−/−, P = 0.0304). (D) Reduced number of PMELs in the epidermis A global defect in priming in our model is unlikely because the shift in the
of Cxcl9−/− mice and a trend toward reduced PMELs in Cxcl10−/− mice com- ratio of CD44lo to CD44hi cells is not present within skin-draining lymph
pared to WT mice (one-way ANOVA P = 0.022 with Tukey’s post-tests: WT nodes. CD44 is important for the survival, activation, directed migration
versus Cxcl9−/−, P = 0.0223). (C and D) Data were pooled from two separate (particularly through basement membranes), and general “fitness” of mem-
experiments (n = 11 WT, 11 Cxcl9−/−, and 8 Cxcl10−/− age-matched male ory CD8+ T cells within peripheral tissues (49). Whether our observations
mice). (E) The ratio of antigen-experienced (CD69+) CD44lo to CD44hi PMELs
represent a direct effect of CXCL10 on CD44 expression in memory cells,
was significantly higher in the epidermis of Cxcl10-deficient mice compared
to WT mice (n = 9 WT, 7 Cxcl9−/−, and 7 Cxcl10−/− pooled from two separate or a secondary effect after mislocalization within the skin, is unknown.
experiments, zero ratios not reported; one-way ANOVA with Tukey’s post- The specific roles CXCL10 plays within the skin in vitiligo are important
tests: WT versus Cxcl10−/−, P = 0.0478). (F) However, this ratio was un- but probably complex and will require further study.
changed in skin-draining lymph nodes, indicating an epidermis-specific Previous studies, including ours, reported that interfering with
defect. IFN-g prevents the onset of depigmentation in various mouse models

www.ScienceTranslationalMedicine.org 12 February 2014 Vol 6 Issue 223 223ra23 6


RESEARCH ARTICLE

of vitiligo, including genetic deficiency or antibody neutralization MATERIALS AND METHODS


(26, 33). However, successful treatment strategies should be able to
reverse established disease through repigmentation, rather than simply Study design
prevent it. Depigmentation in patients with vitiligo can be reversed The overall study design was based on controlled laboratory experi-
upon successful treatment with topical steroids, topical calcineurin in- mentation using ex vivo human tissue samples and a mouse model
hibitors, or nbUVB light therapy (1), and this reversal is primarily for in vivo mechanistic studies. The research objectives at the outset
due to melanocyte regeneration from reservoirs in hair follicles within of the study were to test the hypothesis that IFN-g–inducible chemo-
the lesion (10). CXCL10 antibody treatment induced epidermal repig- kines were responsible for the recruitment of autoreactive T cells to
mentation in mice with established vitiligo and therefore may offer a the skin. This hypothesis was formed on the basis of previously re-
targeted systemic therapeutic option for patients with disease. A ther- ported observations in our mouse model (26). Sample size was deter-
apeutic strategy based on these observations may incorporate existing mined using the approach described by Dell et al. (60). Briefly, each
neutralizing antibodies or chemical inhibitors of IFN-g (50), CXCL10 experiment was powered to detect a difference between group means
(51), or CXCR3 (41, 52), or inhibitors of IFN-g or CXCL10 signaling of twice the observed SD, with a power of 0.8 and a significance level
through chemical inhibition or RNA interference–induced knockdown. of 0.05. Replicate experiments were performed two or three times.
Completed clinical trials have tested either CXCL10 neutralization or Inclusion/exclusion criteria for human patient samples were as follows:
CXCR3 blockade for psoriasis, Crohn’s disease, and ulcerative colitis, Vitiligo lesional skin was obtained from clinical biopsies that revealed
with disappointing results (41). However, unlike vitiligo, these diseases characteristic features of vitiligo, including a prominent T cell infil-
are characterized by the expression of additional cytokines and chemo- trate. Control skin was acquired from tumor excision tips without no-

Downloaded from stm.sciencemag.org on May 3, 2015


kines that likely reflect a more robust inflammatory response with multiple table pathology from donors without vitiligo. Human blood and serum
redundant pathways mediating T cell recruitment to sites of inflammation samples were collected from patients examined by a dermatologist
(53, 54). Therefore, vitiligo may be the optimal autoimmune disease in which (J.E.H.) and diagnosed with vitiligo, as well as from healthy volunteers
to test compounds that target the IFN-g–CXCL10–CXCR3 cytokine axis. with no known autoimmune diseases. Mice used for these studies were
In addition to its therapeutic implications, the ability of CXCL10 on the C57BL/6J (B6) background or a mixed 129 × C57BL/6 background
neutralization to promote repigmentation in our mouse model reveals that had been backcrossed to B6 for more than 10 generations. Mice were
that (i) mouse melanocytes are capable of regeneration as in human randomly selected and assigned to treatment groups as described below.
vitiligo and (ii) CXCL10 is required to prevent regeneration. The abil- Scoring of vitiligo progression in mice was done by a single investigator
ity of melanocytes to quickly regenerate after treatment suggests that who was blinded to the treatments. For repigmentation experiments,
there is active autoimmunity in depigmented skin that can be disrupted scores were quantified objectively with ImageJ without blinding.
with treatment, including neutralization of CXCL10. Exactly how
CXCL10 maintains depigmentation in vitiligo is unknown, although Study subjects
any of the functional roles for CXCL10 described above (localization, The collection of serum and blood from healthy donors and patients
migration velocity, dendritic cell interaction, or tissue retention) may be with vitiligo was approved by the Institutional Review Board (IRB) at
important. A recent study found that antigen-specific CD8+ resident the University of Massachusetts Medical School (UMMS) and the
memory T cells that remain within the vaginal mucosa after viral in- University of British Columbia. All participants gave written informed
fection act primarily as sentinels to detect reinfection (55). In this role, consent before all procedures. Identification and acquisition of FFPE
they produce large amounts of IFN-g upon antigen reencounter, which clinical tissue samples were IRB-approved at the respective institu-
stimulates production of chemokines to recruit additional central mem- tions, and all samples were de-identified before use in experiments.
ory effectors to help control the infection. Therefore, a population of
PMELs may remain within depigmented skin and act as sentinels for Enzyme-linked immunosorbent assays
melanocyte regeneration, producing IFN-g after melanocyte encounter, Human CXCL9, CXCL10, and CXCL11 DuoSet ELISA kits were
which induces chemokines and recruits additional central memory used to analyze patient serum according to the manufacturer’s in-
PMELs to target these cells and prevent repigmentation. structions (R&D Systems). Optical densities were measured with a
Our studies provide the basis for further exploration of CXCL10 as PerkinElmer EnVision 2102 multilabel reader, and 595-nm values were
a treatment target for vitiligo; however, they have potential limitations. subtracted from 450-nm values to calculate concentrations with a four-
There are reports describing elevated IL-17 in vitiligo patient serum as parameter logarithmic standard curve.
well as IL-17+ T cells in lesional skin (56–59), although the functional
significance of these observations is currently unclear. We did not see Mice
increased expression of IL-17 or the IL-17–dependent chemokine CCL20 KRT14-Kitl*4XTG2Bjl (Krt14-Kitl*) mice were a gift from B. J. Longley
in any of our 13 vitiligo patients. It is possible that our patient popu- (University of Wisconsin). PMEL TCR transgenic mice are Thy1.1+
lation reflects one subset of vitiligo with an IFN-g–specific profile and and were obtained from The Jackson Laboratory, stock no. 005023,
that our mouse model reflects this particular subset. There may be B6.Cg Thy1a/CyTg(TcraTcrb)8Rest/J. GFP-PMEL and CFP-PMEL
multiple subtypes of vitiligo based on cytokine expression, although TCR transgenic mice were produced by crossing PMEL transgenic
this has yet to be determined. Nevertheless, our results identify CXCL10 mice with DPEGFP mice, which express GFP in T cells (provided by
as an important mediator of vitiligo pathogenesis in a mouse model U. von Andrian, Harvard Medical School, Boston, MA), or with
of disease. In addition, they reveal that the maintenance of depig- CAG-ECFP [The Jackson Laboratory, stock no. 004218, B6.129(ICR)-
mentation in vitiligo is an active process of immunity that requires Tg(CAG-ECFP)CK6Nagy/J], respectively. CFP-PMEL TCR transgenic
CXCL10. Ultimately, clinical trials will be required to test the therapeutic mice were backcrossed to B6 for a total of 10 generations. Cxcr3−/−
implications of our observations to patients with this devastating disease. PMEL TCR transgenic mice were produced by crossing GFP-PMEL

www.ScienceTranslationalMedicine.org 12 February 2014 Vol 6 Issue 223 223ra23 7


RESEARCH ARTICLE

mice with the B6.129P2-Cxcr3tm1Dgen/J strain (The Jackson Labora- Flow cytometry
tory, stock no. 005796). To develop Cxcl9- and Cxcl10-deficient re- Ears, tails, spleens, and skin-draining lymph nodes were harvested at
cipients, B6.129S4-Cxcl9tm1Jmf (provided by J. M. Farber, National the indicated times. Spleens were disrupted, and the red blood cells
Institute of Allergy and Infectious Diseases, Bethesda, MD) and were lysed with RBC lysis buffer. Ears and tail skin were incubated
B6.129S4-Cxcl10tm1Adl/J (The Jackson Laboratory, stock no. 006087) in skin digest medium [RPMI containing 0.5% deoxyribonuclease
mice were crossed with Krt14-Kitl*. For consistency, the Krt14-Kitl* (DNase) I (Sigma-Aldrich) and liberase TL enzyme blend (0.5 mg/ml)
allele was heterozygous on all mice used in experiments. All mice were (Roche)] and processed with a medimachine (BD Biosciences) as de-
on a C57BL/6J background and were maintained in pathogen-free fa- scribed previously (26). For separation of the dermis and epidermis,
cilities at UMMS, and procedures were approved by the UMMS Insti- tail skin samples were incubated with dispase (2.4 U/ml) (Roche) for
tutional Animal Care and Use Committee and in accordance with the 1 hour at 37°C. Epidermis was removed and mechanically disrupted
National Institutes of Health (NIH) Guide for the Care and Use of with 70-mm cell strainers, and dermis samples were incubated with
Laboratory Animals. collagenase IV (1 mg/ml) with DNase I (0.5 mg/ml) (Sigma-Aldrich)
for 1 hour at 37°C on a shaker. Human peripheral blood mononuclear
Vitiligo induction and chemokine neutralization cells (PBMCs) were isolated with Histopaque-1077 (Sigma-Aldrich)
Vitiligo was induced through adoptive transfer of PMEL CD8+ T cells and washed in PBS. All cells were filtered through a 70-mm mesh be-
as described previously (26). Briefly, PMEL CD8+ T cells were isolated fore analysis. The following antibodies were obtained from BioLegend:
from the spleens of PMEL TCR transgenic mice through negative se- mouse (CD4, CD8b, CD45.2, CD90.1, CXCR3, CD44, CD69, and Fc
lection on microbeads (Miltenyi Biotec) according to the manufacturer’s block) and human (CD3, CD4, CD8, CD19, CD45, and CXCR3).

Downloaded from stm.sciencemag.org on May 3, 2015


instructions. Purified CD8+ T cells (1×106) were injected intravenously The following fluorescent-conjugated pentamers were obtained from
into sublethally irradiated (500 rads 1 day before transfer) Krt14- ProImmune Ltd.: tyrosinase 369–377 (371D) YMDGTMSQV, gp100
Kitl* hosts (12 to 16 weeks of age). Recipient mice also received intra- (PMEL) 209–217 ITDQVPFSV. PBMCs (10 × 106) were allocated per
peritoneal injection of 1 × 106 plaque-forming units of rVV-hPMEL staining condition. To improve the sensitivity of detection, the cells
(N. Restifo, National Cancer Institute, NIH) on the same day of trans- were incubated first at 37°C for 30 min in 50 ml of dasatinib solution
fer. Mice used as controls for vitiligo were sublethally irradiated but (50 nM) (Axon Medchem BV) as previously described (61). Then,
did not receive rVV-hPMEL or PMEL CD8+ T cell transfer. To measure 10 ml of labeled pentamers was added and incubated at room tempera-
the migration efficiency of Cxcr3−/− T cells to the skin within the same ture for 10 min, followed by washing and the addition of secondary
host, equal numbers (0.5 × 106) of CFP+ wild-type T cells and GFP+ antibodies. Pentamer-positive cells were identified by gating on CD45+,
wild-type or Cxcr3−/− T cells were injected intravenously. CD19−, CD3+, CD4−, CD8+ lymphoid cells (fig. S2). The data were col-
CXCL9 and CXCL10 blockade was performed by intraperitoneal in- lected and analyzed with a BD LSR II flow cytometer (BD Biosciences)
jection of 100 mg of either CXCL9-neutralizing (clone 2A6) or CXCL10- and FlowJo (Tree Star Inc.).
neutralizing (clone 1F11) antibodies three times weekly for the duration
of the observation period (3 to 5 weeks). Mice with vitiligo used as Gene expression profiling and validation
treatment controls for chemokine blockade were treated either with Mouse Whole-Genome 6 version 2.0 (Illumina Inc.) was used for ex-
no antibody or with an equal volume of PBS. Vitiligo score was ob- pression profiling of fresh mouse ear skin, and Whole-Genome DASL
jectively quantified by an observer blinded to the experimental groups, HT Assay version 4.0 (Illumina Inc.) was used for expression profiling
using a point scale based on the extent of depigmentation at four of FFPE human samples, according to the manufacturer’s instructions.
easily visible locations, including the ears, nose, rear footpads, and Ear skin from mice with vitiligo was compared to control mice that
tails as described previously (26). Each location was examined, and were irradiated but did not receive PMEL TCR transgenic T cells or
the extent of depigmentation was estimated as a percentage of the VV-gp100. Lesional skin with characteristic features of vitiligo and
anatomic site; both left and right ears and left and right rear footpads a notable T cell infiltrate was identified from five vitiligo patients in a
were estimated together and therefore evaluated as single sites. Points clinical biopsy database and compared to five age- and site-matched
were awarded as follows: no evidence of depigmentation (0%) re- controls, which were obtained from tumor excision tips (University of
ceived a score of 0, >0 to 10% = 1 point, >10 to 25% = 2 points, >25 Pennsylvania, Philadelphia, PA). Fold changes were calculated using
to 75% = 3 points, >75 to <100% = 4 points, and 100% = 5 points. The quantile normalized data (MATLAB 2012b) by dividing the average
“vitiligo score” was the sum of the scores at all four sites, with a intensity of vitiligo samples by the average intensity of controls. Only
maximum score of 20 points. genes significantly detected over background (P < 0.05) in at least one
To induce repigmentation, 7 to 8 weeks after induction of vitiligo, sample were considered.
mice with at least 50% tail depigmentation were randomly assigned to Expression of genes of interest in mouse tissues was confirmed
receive either CXCL10-neutralizing antibody as before or control treat- using quantitative real-time reverse transcription polymerase chain re-
ment (PBS or isotype antibody as indicated) for a total of 8 weeks. action (PCR) on ear skin from an additional five control and six vitiligo
Treatment efficacy was objectively quantified by comparison of the tail mice. Briefly, ears were minced into <5-mm strips, stored in RNAlater
photographs before and after treatment with ImageJ software (NIH). (Ambion) overnight at 4°C, and then transferred to −20°C until RNA
All images were converted to 8-bit black and white, and the brightness was extracted with the RNeasy Plus Mini Kit (Qiagen) according to
threshold was adjusted so that all pigmented areas of the tails were the manufacturer’s instructions. Complementary DNA (cDNA) was gen-
highlighted. Highlighted pixels were quantified, and the mean level erated with iScript cDNA synthesis kit (Bio-Rad). Real-time PCR was
of pigmentation was determined. Fold change over baseline reflects conducted with cDNA and iQ SYBR Green (Bio-Rad) in a Bio-Rad
the mean pigmentation after treatment divided by mean pigmentation iCycler iQ, according to the manufacturer’s recommendations. Mouse
before treatment, and therefore, a score of 1.0 reflects no change. primer sequences are as follows: Mx1, 5′-CGATCAAGACGGC-

www.ScienceTranslationalMedicine.org 12 February 2014 Vol 6 Issue 223 223ra23 8


RESEARCH ARTICLE

CATGAGTC-3′ (sense) and 5′-CTCGTCGGTGTCATCTTCTGC-3′ REFERENCES AND NOTES


(antisense); Cxcl9, 5′-ATCTCCGTTCTTCAGTGTAGCAATG-3′
(sense) and 5′-ACAAATCCCTCAAAGACCTCAAACAG-3′ (antisense); 1. A. Taïeb, M. Picardo, Clinical practice. Vitiligo. N. Engl. J. Med. 360, 160–169 (2009).
2. T. Passeron, J. P. Ortonne, Activation of the unfolded protein response in vitiligo: The
Cxcl10, 5′-AGGGGAGTGATGGAGAGAGG-3′ (sense) and 5′-
missing link? J. Invest. Dermatol. 132, 2502–2504 (2012).
TGAAAGCGTTTAGCCAAAAAAGG-3′ (antisense); and b-actin (Actb), 3. A. A. Shah, A. A. Sinha, Oxidative stress and autoimmune skin disease. Eur. J. Dermatol. 23,
5′-GGCTGTATTCCCCTCCATCG-3′ (sense) and 5′-CCAGTTGGTA- 5–13 (2013).
ACAATGCCATGT-3′ (antisense). Gene expression is reported relative 4. N. C. Laddha, M. Dwivedi, M. S. Mansuri, A. R. Gani, M. Ansarullah, A. V. Ramachandran, S. Dalai,
to the average expression in control mice after normalization to expres- R. Begum, Vitiligo: Interplay between oxidative stress and immune system. Exp. Dermatol. 22,
245–250 (2013).
sion of b-actin. 5. J. G. van den Boorn, C. J. Melief, R. M. Luiten, Monobenzone-induced depigmentation:
Human microarray data were validated with nCounter Analysis Sys- From enzymatic blockade to autoimmunity. Pigment Cell Melanoma Res. 24, 673–679 (2011).
tem (NanoString Technologies Inc.) in a separate series of FFPE sam- 6. S. Toosi, S. J. Orlow, P. Manga, Vitiligo-inducing phenols activate the unfolded protein
ples from an additional three healthy and eight vitiligo skin biopsies response in melanocytes resulting in upregulation of IL6 and IL8. J. Invest. Dermatol.
132, 2601–2609 (2012).
distinct from those used for the Illumina DASL assay. Samples were
7. J. A. Mosenson, A. Zloza, J. Klarquist, A. J. Barfuss, J. A. Guevara-Patino, I. C. Poole, HSP70i is
selected as before, and RNA was isolated by High Pure RNA Paraffin a critical component of the immune response leading to vitiligo. Pigment Cell Melanoma Res.
Kit (Roche) and analyzed by Illumina Bioanalyzer. As per the manu- 25, 88–98 (2012).
facturer’s recommendations, RNA samples were adjusted to input 8. J. A. Mosenson, A. Zloza, J. D. Nieland, E. Garrett-Mayer, J. M. Eby, E. J. Huelsmann,
~100 ng of RNA fragments greater than 300 nucleotides, which was P. Kumar, C. J. Denman, A. T. Lacek, F. J. Kohlhapp, A. Alamiri, T. Hughes, S. D. Bines, H. L. Kaufman,
A. Overbeck, S. Mehrotra, C. Hernandez, M. I. Nishimura, J. A. Guevara-Patino, I. C. Le Poole,
then hybridized to the cartridge. Expression for each sample was nor-

Downloaded from stm.sciencemag.org on May 3, 2015


Mutant HSP70 reverses autoimmune depigmentation in vitiligo. Sci. Transl. Med. 5, 174ra128
malized to 10 internal reference genes, and all data analysis was per- (2013).
formed according to the manufacturer’s recommendations. 9. A. Alikhan, L. M. Felsten, M. Daly, V. Petronic-Rosic, Vitiligo: A comprehensive overview.
Part I. Introduction, epidemiology, quality of life, diagnosis, differential diagnosis, associations,
histopathology, etiology, and work-up. J. Am. Acad. Dermatol. 65, 473–491 (2011).
Immunohistochemistry
10. S. J. Glassman, Vitiligo, reactive oxygen species and T-cells. Clin. Sci. 120, 99–120 (2011).
Immunohistochemical studies were performed on 5-mm sections of 11. K. U. Schallreuter, P. Bahadoran, M. Picardo, A. Slominski, Y. E. Elassiuty, E. H. Kemp,
four FFPE clinical biopsy specimens with characteristic features of vit- C. Giachino, J. B. Liu, R. M. Luiten, T. Lambe, I. C. Le Poole, I. Dammak, H. Onay, M. A. Zmijewski,
iligo and a notable mononuclear infiltrate. Samples were distinct from M. L. Dell’Anna, M. P. Zeegers, R. J. Cornall, R. Paus, J. P. Ortonne, W. Westerhof, Vitiligo
those used for gene expression analyses. Slides were first deparaffinized pathogenesis: Autoimmune disease, genetic defect, excessive reactive oxygen species,
calcium imbalance, or what else? Exp. Dermatol. 17, 139–140 (2008).
and then rehydrated. The sections were blocked for endogenous per- 12. R. A. Spritz, Six decades of vitiligo genetics: Genome-wide studies provide insights into
oxidase activity with an application of Dual Endogenous Block (Dako) autoimmune pathogenesis. J. Invest. Dermatol. 132, 268–273 (2012).
and then incubated with one of the following antibodies: anti-CD3 mouse 13. J. G. van den Boorn, D. Konijnenberg, T. A. Dellemijn, J. P. van der Veen, J. D. Bos, C. J. Melief,
monoclonal antibody, clone F7.2.38 (Dako), at a dilution of 1:200; anti- F. A. Vyth-Dreese, R. M. Luiten, Autoimmune destruction of skin melanocytes by perilesional
T cells from vitiligo patients. J. Invest. Dermatol. 129, 2220–2232 (2009).
CD8 mouse monoclonal antibody, clone C8/144B (Dako), at a dilution 14. G. S. Ogg, P. Rod Dunbar, P. Romero, J. L. Chen, V. Cerundolo, High frequency of skin-homing
of 1:80; or anti-CXCR3 mouse monoclonal antibody, clone 1C6/CXR3 melanocyte-specific cytotoxic T lymphocytes in autoimmune vitiligo. J. Exp. Med. 188,
(BD Pharmingen), at a dilution of 1:100. After a rinse in wash buffer 1203–1208 (1998).
(1× tris-buffered saline containing 0.05% Tween 20), sections were incu- 15. R. van den Wijngaard, A. Wankowicz-Kalinska, C. Le Poole, B. Tigges, W. Westerhof, P. Das,
Local immune response in skin of generalized vitiligo patients. Destruction of melanocytes
bated with the Dako EnVision+ HRP detection reagent, washed, and is associated with the prominent presence of CLA+ T cells at the perilesional site. Lab. Invest.
treated with a solution of diaminobenzidine and hydrogen peroxide 80, 1299–1309 (2000).
(Dako) for 10 min to produce the visible brown pigment. 16. I. C. Le Poole, R. M. van den Wijngaard, W. Westerhof, P. K. Das, Presence of T cells and
macrophages in inflammatory vitiligo skin parallels melanocyte disappearance. Am. J. Pathol.
148, 1219–1228 (1996).
Statistical analyses 17. L. M. Felsten, A. Alikhan, V. Petronic-Rosic, Vitiligo: A comprehensive overview. Part II:
All statistical analyses were performed with GraphPad Prism software. Treatment options and approach to treatment. J. Am. Acad. Dermatol. 65, 493–514
Dual comparisons were made with unpaired Student’s t test (parame- (2011).
tric data, mouse studies) or the Mann-Whitney U test (nonparametric 18. A. Taieb, A. Alomar, M. Böhm, M. L. Dell’anna, A. De Pase, V. Eleftheriadou, K. Ezzedine,
Y. Gauthier, D. J. Gawkrodger, T. Jouary, G. Leone, S. Moretti, L. Nieuweboer-Krobotova,
data, human ELISAs). Groups of three or more were analyzed by
M. J. Olsson, D. Parsad, T. Passeron, A. Tanew, W. van der Veen, N. van Geel, M. Whitton,
ANOVA with Tukey’s post-tests for experiments comparing genotypes A. Wolkerstorfer, M. Picardo; Vitiligo European Task Force (VETF); European Academy of
of mice or with Dunnett’s post-tests for experiments comparing treat- Dermatology and Venereology (EADV); Union Européenne des Médecins Spécialistes
ments to controls. P values < 0.05 were considered significant. (UEMS), Guidelines for the management of vitiligo: The European Dermatology Forum
consensus. Br. J. Dermatol. 168, 5–19 (2013).
19. E. L. Baker, C. I. Coleman, K. M. Reinhart, O. J. Phung, L. Kugelman, W. Chen, C. M. White,
C. M. Mamolo, J. C. Cappelleri, W. L. Baker, Effect of biologic agents on non-PASI out-
SUPPLEMENTARY MATERIALS comes in moderate-to-severe plaque psoriasis: Systematic review and meta-analyses.
Dermatol. Ther. 2, 9 (2012).
www.sciencetranslationalmedicine.org/cgi/content/full/6/223/223ra23/DC1 20. K. M. Alghamdi, H. Khurrum, A. Taieb, K. Ezzedine, Treatment of generalized vitiligo with
Fig. S1. Validation of selected transcripts in humans and mice. anti-TNF-a agents. J. Drugs Dermatol. 11, 534–539 (2012).
Fig. S2. Frequencies of melanocyte antigen–specific T cells in peripheral blood of healthy or 21. K. M. Alghamdi, H. Khurrum, A. Rikabi, Worsening of vitiligo and onset of new psoriasiform
vitiligo patients. dermatitis following treatment with infliximab. J. Cutan. Med. Surg. 15, 280–284 (2011).
Fig. S3. Staining controls for vitiligo biopsy immunohistochemistry. 22. S. Bin Dayel, K. AlGhamdi, Failure of alefacept in the treatment of vitiligo. J. Drugs Dermatol. 12,
Fig. S4. Gating strategy for examining CXCR3+ mouse T cells. 159–161 (2013).
Fig. S5. Cotransfers of wild-type and CXCR3-deficient PMELs. 23. A. Gilhar, T. Pillar, S. Eidelman, A. Etzioni, Vitiligo and idiopathic guttate hypomelanosis.
Fig. S6. Repigmentation after treatment with CXCL10-neutralizing antibody but not isotype Repigmentation of skin following engraftment onto nude mice. Arch. Dermatol. 125,
control antibody. 1363–1366 (1989).
Table S1. Percentages of CXCR3+ and CXCR3− PMELs in vitiligo mice, gated on total CD8+ cells. 24. H. Iizuka, Epidermal turnover time. J. Dermatol. Sci. 8, 215–217 (1994).

www.ScienceTranslationalMedicine.org 12 February 2014 Vol 6 Issue 223 223ra23 9


RESEARCH ARTICLE

25. P. E. Grimes, R. Morris, E. Avaniss-Aghajani, T. Soriano, M. Meraz, A. Metzger, Topical tacrolimus 48. J. Ho, C. C. Kurtz, M. Naganuma, P. B. Ernst, F. Cominelli, J. Rivera-Nieves, A CD8+/CD103high
therapy for vitiligo: Therapeutic responses and skin messenger RNA expression of pro- T cell subset regulates TNF-mediated chronic murine ileitis. J. Immunol. 180, 2573–2580 (2008).
inflammatory cytokines. J. Am. Acad. Dermatol. 51, 52–61 (2004). 49. B. J. Baaten, R. Tinoco, A. T. Chen, L. M. Bradley, Regulation of antigen-experienced T cells:
26. J. E. Harris, T. H. Harris, W. Weninger, E. J. Wherry, C. A. Hunter, L. A. Turka, A mouse model Lessons from the quintessential memory marker CD44. Front. Immunol. 3, 23 (2012).
of vitiligo with focused epidermal depigmentation requires IFN-g for autoreactive CD8+ T-cell 50. D. W. Hommes, T. L. Mikhajlova, S. Stoinov, D. Stimac, B. Vucelic, J. Lonovics, M. Zákuciová,
accumulation in the skin. J. Invest. Dermatol. 132, 1869–1876 (2012). G. D’Haens, G. Van Assche, S. Ba, S. Lee, T. Pearce, Fontolizumab, a humanised anti-interferon g
27. C. Sanda, P. Weitzel, T. Tsukahara, J. Schaley, H. J. Edenberg, M. A. Stephens, J. N. McClintick, antibody, demonstrates safety and clinical activity in patients with moderate to severe Crohn’s
L. M. Blatt, L. Li, L. Brodsky, M. W. Taylor, Differential gene induction by type I and type II disease. Gut 55, 1131–1137 (2006).
interferons and their combination. J. Interferon Cytokine Res. 26, 462–472 (2006). 51. L. Mayer, W. J. Sandborn, Y. Stepanov, K. Geboes, R. Hardi, M. Yellin, X. Tao, L. A. Xu,
28. J. R. Groom, A. D. Luster, CXCR3 ligands: Redundant, collaborative and antagonistic L. Salter-Cid, S. Gujrathi, R. Aranda, A. Y. Luo, Anti-IP-10 antibody (BMS-936557) for ulcerative
functions. Immunol. Cell Biol. 89, 207–215 (2011). colitis: A phase II randomised study. Gut 10.1136/gutjnl-2012-303424 (2013).
29. R. W. McMurray, Adhesion molecules in autoimmune disease. Semin. Arthritis Rheum. 25, 52. C. H. Jenh, M. A. Cox, L. Cui, E. P. Reich, L. Sullivan, S. C. Chen, D. Kinsley, S. Qian, S. H. Kim,
215–233 (1996). S. Rosenblum, J. Kozlowski, J. S. Fine, P. J. Zavodny, D. Lundell, A selective and potent
30. M. Rotondi, A. Falorni, A. De Bellis, S. Laureti, P. Ferruzzi, P. Romagnani, A. Buonamano, CXCR3 antagonist SCH 546738 attenuates the development of autoimmune diseases
E. Lazzeri, C. Crescioli, M. Mannelli, F. Santeusanio, A. Bellastella, M. Serio, Elevated and delays graft rejection. BMC Immunol. 13, 2 (2012).
serum interferon-g-inducible chemokine-10/CXC chemokine ligand-10 in autoimmune primary 53. G. P. Christophi, R. Rong, P. G. Holtzapple, P. T. Massa, S. K. Landas, Immune markers and
adrenal insufficiency and in vitro expression in human adrenal cells primary cultures after stim- differential signaling networks in ulcerative colitis and Crohn’s disease. Inflamm. Bowel Dis.
ulation with proinflammatory cytokines. J. Clin. Endocrinol. Metab. 90, 2357–2363 (2005). 18, 2342–2356 (2012).
31. A. Antonelli, S. M. Ferrari, S. Frascerra, F. Galetta, F. Franzoni, A. Corrado, M. Miccoli, 54. B. J. Nickoloff, H. Xin, F. O. Nestle, J. Z. Qin, The cytokine and chemokine network in pso-
S. Benvenga, A. Paolicchi, E. Ferrannini, P. Fallahi, Circulating chemokine (CXC motif ) riasis. Clin. Dermatol. 25, 568–573 (2007).
ligand (CXCL)9 is increased in aggressive chronic autoimmune thyroiditis, in association 55. J. M. Schenkel, K. A. Fraser, V. Vezys, D. Masopust, Sensing and alarm function of resident
with CXCL10. Cytokine 55, 288–293 (2011). memory CD8+ T cells. Nat. Immunol. 14, 509–513 (2013).

Downloaded from stm.sciencemag.org on May 3, 2015


32. T. Kunisada, S. Z. Lu, H. Yoshida, S. Nishikawa, M. Mizoguchi, S. Hayashi, L. Tyrrell, 56. C. Q. Wang, A. E. Cruz-Inigo, J. Fuentes-Duculan, D. Moussai, N. Gulati, M. Sullivan-Whalen,
D. A. Williams, X. Wang, B. J. Longley, Murine cutaneous mastocytosis and epidermal mel- P. Gilleaudeau, J. A. Cohen, J. G. Krueger, Th17 cells and activated dendritic cells are
anocytosis induced by keratinocyte expression of transgenic stem cell factor. J. Exp. Med. increased in vitiligo lesions. PLOS One 6, e18907 (2011).
187, 1565–1573 (1998). 57. M. A. Elela, R. A. Hegazy, M. M. Fawzy, L. A. Rashed, H. Rasheed, Interleukin 17, interleukin
33. R. K. Gregg, L. Nichols, Y. Chen, B. Lu, V. H. Engelhard, Mechanisms of spatial and temporal 22 and FoxP3 expression in tissue and serum of non-segmental vitiligo: A case- controlled
development of autoimmune vitiligo in tyrosinase-specific TCR transgenic mice. J. Immunol. study on eighty-four patients. Eur. J. Dermatol. 23, 350–355 (2013).
184, 1909–1917 (2010). 58. M. K. Tembhre, V. K. Sharma, A. Sharma, P. Chattopadhyay, S. Gupta, T helper and regulatory T
34. J. R. Groom, J. Richmond, T. T. Murooka, E. W. Sorensen, J. H. Sung, K. Bankert, U. H. von Andrian, cell cytokine profile in active, stable and narrow band ultraviolet B treated generalized vitiligo.
J. J. Moon, T. R. Mempel, A. D. Luster, CXCR3 chemokine receptor-ligand interactions in the Clin. Chim. Acta 424, 27–32 (2013).
lymph node optimize CD4+ T helper 1 cell differentiation. Immunity 37, 1091–1103 (2012). 59. R. Khan, S. Gupta, A. Sharma, Circulatory levels of T-cell cytokines (interleukin [IL]-2, IL-4, IL-17, and
35. S. N. Mueller, T. Gebhardt, F. R. Carbone, W. R. Heath, Memory T cell subsets, migration transforming growth factor-b) in patients with vitiligo. J. Am. Acad. Dermatol. 66, 510–511 (2012).
patterns, and tissue residence. Annu. Rev. Immunol. 31, 137–161 (2013). 60. R. B. Dell, S. Holleran, R. Ramakrishnan, Sample size determination. ILAR J. 43, 207–213 (2002).
36. J. H. Sung, H. Zhang, E. A. Moseman, D. Alvarez, M. Iannacone, S. E. Henrickson, J. C. de la Torre, 61. A. Lissina, K. Ladell, A. Skowera, M. Clement, E. Edwards, R. Seggewiss, H. A. van den Berg,
J. R. Groom, A. D. Luster, U. H. von Andrian, Chemokine guidance of central memory T cells is E. Gostick, K. Gallagher, E. Jones, J. J. Melenhorst, A. J. Godkin, M. Peakman, D. A. Price, A. K. Sewell,
critical for antiviral recall responses in lymph nodes. Cell 150, 1249–1263 (2012). L. Wooldridge, Protein kinase inhibitors substantially improve the physical detection of T-cells
37. W. Kastenmüller, M. Brandes, Z. Wang, J. Herz, J. G. Egen, R. N. Germain, Peripheral with peptide-MHC tetramers. J. Immunol. Methods 340, 11–24 (2009).
prepositioning and local CXCL9 chemokine-mediated guidance orchestrate rapid memory 62. R. Edgar, M. Domrachev, A. E. Lash, Gene Expression Omnibus: NCBI gene expression and
CD8+ T cell responses in the lymph node. Immunity 38, 502–513 (2013). hybridization array data repository. Nucleic Acids Res. 30, 207–210 (2002).
38. A. Wańkowicz-Kalińska, R. M. van den Wijngaard, B. J. Tigges, W. Westerhof, G. S. Ogg,
V. Cerundolo, W. J. Storkus, P. K. Das, Immunopolarization of CD4+ and CD8+ T cells to type-1–like Acknowledgments: We thank clinic patients (of J.E.H. and Y.Z.) for donating blood, B. J. Longley
is associated with melanocyte loss in human vitiligo. Lab. Invest. 83, 683–695 (2003). for Krt14-Kitl* mice, U. von Andrian for DPEGFP mice, J. M. Farber for Cxcl9-deficient mice, N. Restifo
39. D. Wong, B. Kea, R. Pesich, B. W. Higgs, W. Zhu, P. Brown, Y. Yao, D. Fiorentino, Interferon for recombinant vaccinia virus, M. Frisoli and K. Essien for technical assistance, and M. Brehm, D. Greiner,
and biologic signatures in dermatomyositis skin: Specificity and heterogeneity across and C. Hartigan for critical support in collecting blood for ELISA and flow analysis. Funding: Supported
diseases. PLOS One 7, e29161 (2012). by the National Institute of Arthritis and Musculoskeletal and Skin Diseases, part of the NIH, under Award
40. J. Wenzel, T. Tüting, An IFN-associated cytotoxic cellular immune response against viral, Number AR061437, and research grants from the Charles H. Hood Foundation, Vitiligo Research
self-, or tumor antigens is a common pathogenetic feature in “interface dermatitis”. Foundation, and Dermatology Foundation (to J.E.H.). The University of Massachusetts Center for
J. Invest. Dermatol. 128, 2392–2402 (2008). Clinical Research was responsible for blood and serum collection and is supported by NIH Clinical
41. M. Wijtmans, D. Verzijl, R. Leurs, I. J. de Esch, M. J. Smit, Towards small-molecule CXCR3 and Translational Sciences Award UL1TR000161. The University of Pennsylvania Skin Disease Re-
ligands with clinical potential. ChemMedChem 3, 861–872 (2008). search Center Core was responsible for identification and processing of human tissue samples for
42. J. G. Hyun, G. Lee, J. B. Brown, G. R. Grimm, Y. Tang, N. Mittal, R. Dirisina, Z. Zhang, gene expression analysis and is supported by NIH grant 5-P30-AR057217-03. The Wistar Institute
J. P. Fryer, J. V. Weinstock, A. D. Luster, T. A. Barrett, Anti-interferon-inducible chemokine, Genomics Core performed Illumina gene expression analyses and is supported by NIH grant P30
CXCL10, reduces colitis by impairing T helper-1 induction and recruitment in mice. CA010815. Flow cytometry equipment used for this study is maintained by the UMMS and the Uni-
Inflamm. Bowel Dis. 11, 799–805 (2005). versity of Pennsylvania Flow Cytometry Core Facility. Author contributions: J.E.H., A.D.L., C.A.H., and
43. J. Menke, G. C. Zeller, E. Kikawada, T. K. Means, X. R. Huang, H. Y. Lan, B. Lu, J. Farber, A.D. designed the study. M.R., P.A., J.M.R., T.H.H., and K.D. performed the experiments. M.-W.S., Y.Z.,
A. D. Luster, V. R. Kelley, CXCL9, but not CXCL10, promotes CXCR3-dependent immune- A.D., and J.E.H. provided patient and patient sample management. M.R., J.M.R., K.D., and J.E.H.
mediated kidney disease. J. Am. Soc. Nephrol. 19, 1177–1189 (2008). drafted the manuscript, and A.D.L., C.A.H., J.E.H., J.M.R., and T.H.H. critically revised the manuscript.
44. G. S. Campanella, A. M. Tager, J. K. El Khoury, S. Y. Thomas, T. A. Abrazinski, L. A. Manice, Competing interests: The authors declare that they have no competing interests. Data and
R. A. Colvin, A. D. Luster, Chemokine receptor CXCR3 and its ligands CXCL9 and CXCL10 are materials availability: The data discussed in this publication have been deposited in National Center
required for the development of murine cerebral malaria. Proc. Natl. Acad. Sci. U.S.A. 105, for Biotechnology Information’s Gene Expression Omnibus (GEO) (62) and are accessible through GEO
4814–4819 (2008). Series accession number GSE53148 (http://www.ncbi.nlm.nih.gov/geo/query/acc.cgi?acc=GSE53148).
45. M. Thapa, R. S. Welner, R. Pelayo, D. J. Carr, CXCL9 and CXCL10 expression are critical for control
of genital herpes simplex virus type 2 infection through mobilization of HSV-specific CTL and NK Submitted 15 October 2013
cells to the nervous system. J. Immunol. 180, 1098–1106 (2008). Accepted 23 December 2013
46. T. H. Harris, E. J. Banigan, D. A. Christian, C. Konradt, E. D. Tait Wojno, K. Norose, E. H. Wilson, Published 12 February 2014
B. John, W. Weninger, A. D. Luster, A. J. Liu, C. A. Hunter, Generalized Lévy walks and the 10.1126/scitranslmed.3007811
role of chemokines in migration of effector CD8+ T cells. Nature 486, 545–548 (2012).
47. H. Yoneyama, S. Narumi, Y. Zhang, M. Murai, M. Baggiolini, A. Lanzavecchia, T. Ichida, Citation: M. Rashighi, P. Agarwal, J. M. Richmond, T. H. Harris, K. Dresser, M.-W. Su, Y. Zhou,
H. Asakura, K. Matsushima, Pivotal role of dendritic cell–derived CXCL10 in the retention of A. Deng, C. A. Hunter, A. D. Luster, J. E. Harris, CXCL10 is critical for the progression and
T helper cell 1 lymphocytes in secondary lymph nodes. J. Exp. Med. 195, 1257–1266 (2002). maintenance of depigmentation in a mouse model of vitiligo. Sci. Transl. Med. 6, 223ra23 (2014).

www.ScienceTranslationalMedicine.org 12 February 2014 Vol 6 Issue 223 223ra23 10


CXCL10 Is Critical for the Progression and Maintenance of
Depigmentation in a Mouse Model of Vitiligo
Mehdi Rashighi et al.
Sci Transl Med 6, 223ra23 (2014);
DOI: 10.1126/scitranslmed.3007811

Editor's Summary

New Skin in the Game of Vitiligo Therapy

The immune system is tasked with protecting the body from invading pathogens. Yet, sometimes, immune cells
themselves attack the tissues they are supposed to protect. One such autoimmune disease is vitiligo, where immune
cells are thought to attack melanocytes−−the pigment-producing cells in the skin. Individuals with vitiligo have
depigmented areas in their skin, which is disfiguring and also increases the risk of skin damage. Now, Rashighi et al.
suggest that blocking the chemokine CXCL10 may restore pigmentation in patients with vitiligo.

The authors examined gene expression in lesional skin from vitiligo patients and found an interferon-γ−specific
signature, including differential expression of the chemokine CXCL10. They found that CXCL10 was up-regulated in
vitiligo patients; its receptor CXCR3 was up-regulated in T cells from these patients as well. The authors then looked

Downloaded from stm.sciencemag.org on May 3, 2015


in a mouse model of vitiligo to determine the functional relevance of this observation. Mice with CXCR3-deficient T
cells developed a much less severe form of vitiligo, as did mice lacking CXCL10 or treated with a
CXCL10-neutralizing antibody. What's more, this CXCL10-neutralizing antibody resulted in repigmentation in mice
with already established vitiligo lesions. These data suggest that CXCL10 neutralization should be considered as a
potential treatment for vitiligo.

A complete electronic version of this article and other services, including high-resolution figures,
can be found at:
http://stm.sciencemag.org/content/6/223/223ra23.full.html
Supplementary Material can be found in the online version of this article at:
http://stm.sciencemag.org/content/suppl/2014/02/10/6.223.223ra23.DC1.html
Related Resources for this article can be found online at:
http://stm.sciencemag.org/content/scitransmed/5/174/174ra28.full.html
http://www.sciencemag.org/content/sci/346/6212/932.full.html
http://www.sciencemag.org/content/sci/346/6212/934.full.html
http://www.sciencemag.org/content/sci/346/6212/937.full.html
http://www.sciencemag.org/content/sci/346/6212/941.full.html
http://www.sciencemag.org/content/sci/346/6212/945.full.html
http://www.sciencemag.org/content/sci/346/6212/950.full.html
http://www.sciencemag.org/content/sci/346/6212/954.full.html
Information about obtaining reprints of this article or about obtaining permission to reproduce this
article in whole or in part can be found at:
http://www.sciencemag.org/about/permissions.dtl

Science Translational Medicine (print ISSN 1946-6234; online ISSN 1946-6242) is published weekly, except the
last week in December, by the American Association for the Advancement of Science, 1200 New York Avenue
NW, Washington, DC 20005. Copyright 2014 by the American Association for the Advancement of Science; all
rights reserved. The title Science Translational Medicine is a registered trademark of AAAS.

You might also like