You are on page 1of 42

PHYSIOLOGICAL REVIEWS

Vol. 80, No. 3, July 2000


Printed in U.S.A.

Corticotropin Releasing Hormone and Proopiomelanocortin


Involvement in the Cutaneous Response to Stress
ANDRZEJ SLOMINSKI, JACOBO WORTSMAN, THOMAS LUGER, RALF PAUS, AND SAMUEL SOLOMON

Department of Pathology, Loyola University Medical Center, Maywood; Department of Medicine, Southern
Illinois University, Springfield, Illinois; Department of Dermatology, Ludwig Boltzmann Institute of Cell
Biology and Immunobiology of the Skin, University of Munster, Munster; Department of Dermatology,
University Hospital Eppendorf, University of Hamburg, Hamburg, Germany; and Department of Medicine
and Biochemistry, McGill University and Royal Victoria Hospital, Montreal, Quebec, Canada

I.Physiology of the Systemic Response to Stress: Role of the Hypothalamic-Pituitary-Adrenal Axis 980
II.The Skin as a Shock Organ for Environmental Stresses 980
III.Cutaneous Response to Stress: Local Neuroendocrine Signals 981
IV. Corticotropin Releasing Hormone 981
A. CRH expression: intracranial 981
B. CRH expression: peripheral 982
C. CRH expression: skin 983
V. Corticotropin Releasing Hormone Receptors 984
A. CRH receptors: structure and function 984
B. CRH receptors: skin 985
VI. Proopiomelanocortin 988
A. POMC expression: intracranial 988
B. POMC expression: peripheral 990
C. POMC expression: skin 990
VII. Skin as a Target for Proopiomelanocortin Peptides 997
A. Melanocortin receptors 997
B. Opioid receptors 1000
C. Phenotypic effects of POMC peptides 1000
VIII. Regulation of the Cutaneous Corticotropin Releasing Hormone-Proopiomelanocortin System 1005
A. UVR 1005
B. Cytokines 1005
C. Hair cycle 1005
D. Disease states 1005
E. Cellular metabolism mediators 1005
F. Glucocorticoids 1006
IX. Physiological Significance of Cutaneous Production of Corticotropin Releasing Hormone
and Proopiomelanocortin 1006
X. Hypothesis on an Organized System Mediating Skin Responses to Stress 1006
A. Hypothesis 1006
B. Background to the organization and function of the SSRS 1007
C. Functional organization of SSRS 1008

Slominski, Andrzej, Jacobo Wortsman, Thomas Luger, Ralf Paus, and Samuel Solomon. Corticotropin
Releasing Hormone and Proopiomelanocortin Involvement in the Cutaneous Response to Stress. Physiol Rev 80:
979 –1020, 2000.—The skin is a known target organ for the proopiomelanocortin (POMC)-derived neuropeptides
␣-melanocyte stimulating hormone (␣-MSH), ␤-endorphin, and ACTH and also a source of these peptides. Skin
expression levels of the POMC gene and POMC/corticotropin releasing hormone (CRH) peptides are not static but
are determined by such factors as the physiological changes associated with hair cycle (highest in anagen phase),
ultraviolet radiation (UVR) exposure, immune cytokine release, or the presence of cutaneous pathology. Among the
cytokines, the proinflammatory interleukin-1 produces important upregulation of cutaneous levels of POMC mRNA,
POMC peptides, and MSH receptors; UVR also stimulates expression of all the components of the CRH/POMC
system including expression of the corresponding receptors. Molecular characterization of the cutaneous POMC

www.physrev.physiology.org 0031-9333/00 $15.00 Copyright © 2000 the American Physiological Society 979

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


980 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

gene shows mRNA forms similar to those found in the pituitary, which are expressed together with shorter variants.
The receptors for POMC peptides expressed in the skin are functional and include MC1, MC5 and ␮-opiate, although
most predominant are those of the MC1 class recognizing MSH and ACTH. Receptors for CRH are also present in
the skin. Because expression of, for example, the MC1 receptor is stimulated in a similar dose-dependent manner
by UVR, cytokines, MSH peptides or melanin precursors, actions of the ligand peptides represent a stochastic
(predictable) nonspecific response to environmental/endogenous stresses. The powerful effects of POMC peptides
and probably CRH on the skin pigmentary, immune, and adnexal systems are consistent with stress-neutralizing
activity addressed at maintaining skin integrity to restrict disruptions of internal homeostasis. Hence, cutaneous
expression of the CRH/POMC system is highly organized, encoding mediators and receptors similar to the hypo-
thalamic-pituitary-adrenal (HPA) axis. This CRH/POMC skin system appears to generate a function analogous to the
HPA axis, that in the skin is expressed as a highly localized response which neutralizes noxious stimuli and attendant
immune reactions.

I. PHYSIOLOGY OF THE SYSTEMIC RESPONSE sors such as bursts of radiation (solar, thermal),
TO STRESS: ROLE OF THE HYPOTHALAMIC- mechanical energy, or chemical and biological insults.
PITUITARY-ADRENAL AXIS Because of its functional domains and structural diver-
sity, the skin must have a constitutive mechanism for
The vertebrate brain is endowed with the functional
control of the endocrine system, which proceeds through
highly organized structures. This complex neuroendo-
crine function involves a myriad of pathways and humoral
mediators that are characteristically activated in response
to external (environmental) or internal changes sensed as
stressful. In chronic sustained stresses, the humorally
mediated involvement of the hypothalamic-pituitary-adre-
nal (HPA) axis is most prominent. Activation of this path-
way by the stress-sensoring central circuits or the central
action of proinflammatory cytokines proceeds through
the hypothalamic production and release of corticotropin
releasing hormone (CRH), which stimulates pituitary
CRH receptors (5, 27, 110, 118, 232, 313, 421, 430) (Fig. 1).
CRH enhances the production and secretion of the ante-
rior pituitary-derived propiomelanocortin (POMC) pep-
tides melanocyte stimulating hormone (MSH), ACTH, and
endorphin (17, 110, 208, 267, 268) (Fig. 1). Upon release
into the systemic circulation, ACTH reaches the adrenal
gland and activates the MC2 receptors (MC2-R), inducing
thereby the production and secretion of corticosterone
(rodents) or cortisol (humans). These are powerful anti-
inflammatory factors that counteract the effect of stress
and buffer tissue damage. Moreover, the same steroids act
to terminate the stress response by interacting directly
with central nervous system (CNS) or anterior pituitary
receptors to attenuate CRH and POMC peptide produc-
tion (110, 444). The underlying purpose of this adaptive
response is the stabilization and restoration of general
homeostasis (27, 42, 110, 208, 232, 267, 268, 415).

II. THE SKIN AS A SHOCK ORGAN FOR


FIG. 1. Systemic response to stress. The systemic response to stress
ENVIRONMENTAL STRESSES is activated by environmental/dyshomeostatic signals and biological
insults and is sensed by the central nervous system and the immune
The skin, the largest body organ, is strategically lo- system, respectively. The main mediator of the response is the hypotha-
lamic-pituitary-adrenal axis (HPA). CRH, corticotropin releasing hor-
cated as a barrier between the external and internal en- mone; POMC, proopiomelanocortin; ACTH, adrenocorticotropic hor-
vironments, being permanently exposed to noxious stres- mone; CA, catecholamines.

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 981

dealing with those stressors while cellular/tissue damage


is still localized and of low magnitude, i.e., before trigger-
ing the systemic response. Ideally, such a defense mech-
anism should possess stress-sensoring capability of very
high sensitivity, because of the relatively low intensity of
the continuous pleiotropic physicochemical and biologi-
cal insults affecting the skin. Thus the skin stress re-
sponse mechanism may be envisioned as efficiently re-
cording fluctuating environmental information; once a
critical threshold is reached, this would trigger an orga-
nized biological response. Hence, this cutaneous stress
response mechanism must be efficient, self-regulated in
intensity and field of activity, and endowed with the ca-
pability of differentiating environmental noise from bio-
logically relevant signals (365, 376, 378). Those properties FIG. 3. Chemical mediators and postulated pathways of the func-

imply continuous recognition and integration of appropri- tional arm of the skin CRH/POMC system. URC, urocortin; CRH-Rs, CRH
receptors; Rs, catecholamine receptors; MSH, melanocyte stimulating
ate signals, fast response to activation, and high degree of hormone; ␤-END, ␤-endorphin; MC-Rs, melanocortin receptors; END-R,
specificity, all to rapidly reestablish tissue homeostasis endorphin receptors; 3?, possible stimulatory effect.
(Fig. 2). Desirable responses could involve, for example,
the stimulation of local biosynthetic pathways for the 445). It has been further shown that the skin has the intrinsic
manufacturing of buffering molecules to counteract the capability to actually produce POMC as well as CRH pep-
damaging effect of physical, biological, or chemical in- tides and to also express the corresponding receptors. Cu-
sults (Fig. 3) (365, 376, 378, 382). taneous modulation of CRH and POMC production could be
mediated by proinflammatory cytokines, as is the case at the
systemic level. Thus the skin expresses an equivalent of the
III. CUTANEOUS RESPONSE TO STRESS:
HPA axis that may acts as a cutaneous defense system,
LOCAL NEUROENDOCRINE SIGNALS
operating as coordinator and executor of local responses to
stress (Figs. 2 and 3) (362, 365, 376, 378).
Whereas spatially and temporarily the responses to
Against the background above, we review evidence
stress of skin and CNS may be totally dissociated, their
documenting production and regulation of POMC and CRH
similarity in functional relevance raises the possibility of
peptides, with expression of the corresponding receptors in
shared mediators. In this regard, it is known that mamma-
mammalian skin. Cutaneous mechanism regulating CRH
lian skin, a well-characterized target for the POMC-derived
and POMC gene expression are compared with their central
ACTH and MSH peptides, does contain POMC (220, 376, 438,
homologs, and the function of local CRH and POMC derived
peptides is discussed within the context of the skin response
to stress. Information is provided on the experimental char-
acterization of receptors for CRH and POMC peptides in
skin cells, regulation of their expression, and definition of
signal transduction pathways. Potential interactions within
the epidermal and follicular keratinocyte compartments oc-
curring between the cutaneous equivalent of the hypotha-
lamic-pituitary axis and the skin immune and pigmentary
systems are analyzed against experimental data and clinical
information already available. This review ends by setting
the stage for future research, basic and clinical, on cutane-
ous mediators in response to physical, chemical, or biolog-
ical stressors.

IV. CORTICOTROPIN RELEASING HORMONE

A. CRH Expression: Intracranial


FIG. 2. Cutaneous response to stress. The CRH/POMC system is the
integrator and coordinator of the local response to environmental and CRH, the most proximal element of the HPA axis, has
dyshomeostatic (internal) stimuli. been already sequenced and its gene cloned (348, 413,

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


982 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

427). Most recently, a family of CRH-related peptides has and, through neurons projecting to the pituitary, CRH is
been identified in mammals, which shares high homology involved in osmotic regulation not connected with stress
with CRH, and includes urocortin, amphibian sauvagine, (5, 267). In anterior pituitary corticotrophs, CRH regulates
and fish urotensin I (97, 436). The corresponding genes POMC gene activity and production/release of ACTH,
have also been cloned (344, 348, 413, 454). ␤-endorphin, and MSH peptides (5, 166, 267, 268). In
The CRH gene is composed of two exons separated addition to the hypothalamus, CRH is produced in other
by an intron (344, 348, 413). The first exon encodes most areas of the brain (5, 166). CRH production has even been
of the 5⬘-untranslated region in the mRNA, while the shown in the anterior pituitary, where it may act as a
second exon contains the prohormone sequence and the paracrine stimulator of POMC production, as a growth
3⬘-untranslated region. CRH transcripts in the rodent and factor for corticotrophs, and as a modulator of expression
human brain are ⬃1.4 and 1.5 kb long, respectively (348, of the CRH-R1 gene (132, 239, 309, 413, 433). CRH is also
413). The 3⬘-flanking region contains four polyadenylation
involved in functional modulation of the immune system
addition signals (AATAAA). The 5⬘-flanking region con-
(27, 180, 268a, 331, 392, 394, 418, 430), the reproductive
tains DNA sequences responsible for tissue specific ex-
system (100, 267, 297, 301), and the cardiovascular system
pression, second messenger binding, and glucocorticoid
(19, 112, 267, 316, 320), and it acts as a major catabolic
regulation sites (166, 267, 268). Translation of exon 2
generates the 196-amino acid (aa)-long prepro-CRH, in peptide in the hypothalamus, inhibiting food intake, in-
which the first 26 aa represent the signal peptide, cleaved creasing energy expenditure, and producing sustained
in the rough endoplasmic reticulum to generate pro-CRH- weight loss (165).
(27O196) (49, 166, 294, 295). Unmodified pro-CRH-
(27O196) has 18 kDa molecular mass, that increases to 23
B. CRH Expression: Peripheral
kDa after posttranslational modifications (294, 295). En-
doproteolytic processing of pro-CRH within the trans-
Golgi-network and secretory granules generates the final The CRH gene is widely expressed in extracranial
41-aa-long (4.7 kDa) CRH peptide (166, 267, 268, 294, 295). tissues but at levels much lower than in hypothalamus.
Additional intermediates may include the 16-kDa NH2- Expression of the gene has been detected in endome-
terminal pro-CRH, pro-CRH-(125O149; 8 kDa), and pro- trium, placenta, uterus, ovary, testes, spleen, immune sys-
CRH-(125O151; 3 kDa). The enzymes involved in this tem, pancreas, liver, stomach, small and large bowel,
process are the convertases PC1 and PC2 (49). adrenal gland, thyroid gland, and skin (13, 27, 46, 87, 94,
Expression of the CRH gene is controlled by cAMP- 100, 180, 184, 195, 253, 267, 268, 268a, 298 –303, 305, 318,
dependent protein kinase A (PKA), calcium/calmodulin- 321, 335, 344, 355, 359 –361, 394, 398, 418, 430, 432, 456).
dependent protein kinase and diacylglycerol-dependent Peripheral processing of pro-CRH into CRH appears to be
protein kinase C (PKC) pathways (4, 98, 166, 267, 268, similar at the peripheral and central levels, as shown in
430). In addition, transcription factors associated with placenta, endometrium, uterus, and immune system (49,
cytokine signaling can also activate the CRH promoter 297, 301, 398, 456). In addition to its regulation by similar
(395). A list of stimulators of CRH production at the factors as in the brain (100, 166, 267, 268, 297, 299, 301,
central level includes the neurotransmitters serotonin, 305, 332, 416, 430, 456), CRH production in peripheral
acetylcholine, histamine, norepineprine, and epinephrine tissue is also enhanced by prostaglandins, epidermal
(166, 192, 199, 261, 267, 268); the neuropeptides arginine growth factor, and platelet growth factor and decreased
vasopressin (AVP), ANG II, neuropeptide Y, cholecysto-
by nitric oxide (NO) and progesterone (297, 299, 301, 305,
kinin, activin, and enkephalin; the cytokines interleukin
319, 456). Accumulating evidence indicates that CRH pro-
(IL)-1, IL-6, and tumor necrosis factor (TNF)-␣, which can
duced in uterus and placenta may play an important role
also stimulate CRH production (27, 50, 166, 267, 268, 305,
in the normal progression of pregnancy (235, 297, 301,
395, 418, 420, 421); and leptin (165). Among the negative
regulators of CRH production, the most important are 456). CRH is a potent immunomodulator; depending on
glucocorticoids; estrogens and GABA also share this ac- cellular target, CRH can inhibit or stimulate local immune
tivity (166, 267, 268, 305, 395, 418). In addition, secretion function (9, 142, 195, 268a, 305, 334, 418, 421, 430). CRH
of CRH is inhibited by dynorphin, substance P, somatosta- can modify vascular functions (19, 116, 267, 316, 320) and
tin, and galanin (166). also act as a local growth factor (239, 245, 357, 379, 417,
CRH is produced predominantly in the paraventricu- 433). Expression of the CRH-related urocortin gene with
lar nucleus (PVN) of the hypothalamus and delivered into production of urocortin peptide has also been docu-
portal capillaries converging in the anterior lobe of the mented in peripheral tissues such as placenta, uterus,
pituitary (5, 166, 267, 268, 418). In addition, autonomic immune system, stomach, small and large bowel, pan-
neurons of the PVN projecting to the brain stem and creas, adrenal gland, testis, heart, and skin (18, 262, 263,
spinal cord supply CRH to the sympathoadrenal system 298, 380).

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 983

C. CRH Expression: Skin (321). Corresponding serum concentrations of CRH-IR


were 8.6 fmol/ml during anagen III and 5.6 fmol/ml in
1. Rodent telogen (321). Surprisingly, however, the corresponding
CRH mRNA was consistently below the limit of RT-PCR
CRH was first detected in the skin by immunocyto-
chemistry as CRH immunoreactivity (CRH-IR) in the detectability in any of the different phases of hair cycle
C57BL6 mouse, in the pilosebaceous unit of the hair (telogen, anagen, and catagen) (360). Because CRH-IR is
follicle and epidermis (360). CRH-IR was further localized found in the nerve bundles and perifollicular neural net-
to keratinocytes of the basal epidermis, the outer root work B throughout the entire hair cycle, and CRH con-
sheath (ORS), and the matrix region of developing hair centrations are higher in tissue than serum (321), we have
follicles (321). In hair follicles, the highest intensity of proposed that in the mice CRH is imported into the skin,
stain was seen during the anagen IV/VI stages and the entering through descending (afferent) nerves. Such
lowest levels in catagen and telogen skin (321). CRH-IR transport would provide a mechanism to precisely regu-
was also found in the nerve bundles and perifollicular late domains of local CRH-dependent POMC production
neural network B throughout the entire hair cycle (321). A (321, 357, 360).
schematic drawing of hair cycle coupled CRH and CRH-R The lack of expression of the CRH gene in skin was
expression in murine skin is presented in Figure 4. further confirmed by the results of RT-PCR assays re-
Actual identification of CRH was accomplished with peated with larger number of cycles and followed by
two separate techniques based on reverse-phase HPLC Southern blot hybridization to CRH cDNA; again, these
combined with CRH RIA (321, 357). Tissue levels of CRH tests failed to show CRH mRNA in mouse skin, despite its
measured by RIA showed hair cycle-dependent fluctua- presence in brain control (104). Thus the abundance of
tion being highest in anagen II/IV skin (67 fmol/g wet wt) CRH-IR in the skin and its nerve bundles in the absence of
and lowest in catagen and telogen skin (36 fmol/g wet wt) cutaneous CRH gene expression imply that at least in the

FIG. 4. Expression of CRH and CRH-R1 antigens throughout hair growth cycle in mice. CRF, CRH; APM, arector pili
muscle; DP, dermal papila; E, epidermis; FBN, follicular neural network B; HS, hair shaft; IRS and ORS, inner and outer
root sheaths, respectively; Mel, melanocytes; SG, sebaceous gland. [From Roloff et al. (321), with permission from the
FASEB Journal.]

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


984 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

mice, CRH may reach the skin through descending nerves (73, 77, 78, 190, 194, 207, 217, 242, 296, 391, 426, 437). Both
that target well-defined compartments (360). An alterna- receptor groups share high sequence homology, ⬃70%,
tive explanation, that skin cells could express a hitherto and belong to the family of seven transmembrane seg-
undetected CRH variant or a related CRH gene, is under- ments proteins coupled to the Gs signaling system.
going current testing. In this regard, immunoreactivity CRH-R1 and CRH-R2 differ in pharmacological ligand pro-
corresponding to the related peptide urocortin was de- file and predominant tissue distribution; e.g., CRH-R1 has
tected at varying levels in mouse skin; the highest con- predominantly been found in the pituitary and different
centrations were observed in telogen and the lowest in regions of the brain (5, 74, 78, 289, 310, 437), whereas
late anagen VI skin (380). CHR-R2 has been isolated from brain, heart, and skeletal
muscle (73, 190, 194, 207, 217, 296, 310). Alternatively
2. Human spliced forms have also been identified, e.g., CRH-R1
The CRH gene, although not expressed in mice skin, (variant ␤) that contains 29 additional aa in the first
is nevertheless clearly demonstrable in human skin by intracellular loop and a variant c having a deletion of 120
RT-PCR amplification of the predicted 413-bp transcript bp coding for 40 aa from the extracellular NH2-terminal
representative of the CRH exon 2 (355, 359, 361). In the domain (78, 388). Three CRH-R2 spliced variants ␣, ␤, and
latter experiments, the sources of human skin RNA tested ␥ that differ in their NH2-terminal domains and anatomi-
were biopsy specimens of scalp, compound melanocytic cal distribution have been identified (190, 194, 207, 217,
nevus and basal cell carcinoma, and cultured cell lines of 242, 296, 391, 426). In humans CRH-R2␣ and CRH-R2␤ are
normal and malignant melanocytes and squamous cell expressed in peripheral tissues and in brain, whereas in
carcinoma (355, 359, 361). Melanoma and squamous cell rodents, CRH-R2␣ is found predominantly in brain and
carcinoma cells showed a CRH mRNA transcript 1.5 kb CRH-R2␤ in the periphery (5, 74, 388). So far, CRH-R2␥
long by Northern blot hybridization (361). CRH peptides variant has been only detected in brain (194). Human and
were also detected in facial skin, cultured human mela- rat CRH-R1 genes contain 14 and 13 exons, respectively
nocytes, HaCaT keratinocytes, squamous cell carcinoma, (330, 419), whereas the human CRH-R2␣ gene contains 12
and melanoma cells after RP-HPLC separation followed exons (207). In addition to CRH receptors, CRH binding
by monitoring of the eluted fractions with specific anti- proteins of 322 aa have been characterized; these are
CRH RIA (355, 361). Most recently, expression of CRH partially associated with plasma membranes or circulate
gene was found in a large panel of cultured melanocytes, in the blood (74, 185, 218, 293, 307). CRH binding proteins
nevocytes, and melanoma cells, with the highest intensity have high affinity for CRH and are postulated to inactivate
in malignant melanocytes (121). Indirect immunofluores- extracellular CRH, thus preventing its interaction with
cence studies localized the CRH antigen to keratinocytes receptors. This would be the mechanism for neutraliza-
of the epidermis and hair follicle, dermal blood vessels, tion of the large amounts of circulating CRH present
skeletal muscle, and nerve bundles of human scalp (357). during pregnancy (74, 185, 218, 235, 293, 300, 307).
Expression of the related urocortin gene and production The CRH signal is transduced into cAMP and calcium-
of the corresponding urocortin peptide was also shown in activated metabolic pathways via interaction with the CRH-
human skin (380). R (5, 74, 100, 267, 268, 388). Activation of adenylate cyclase
In normal melanocytes, production of CRH peptide induces production of cAMP, and subsequently of PKA-
could be stimulated with ultraviolet radiation B (UV-B; dependent pathways. CRH-R activation of phospholipase C
wavelength 290 –320 nm) (355); and in melanoma and induces production of inositol trisphosphate (IP3), which
squamous cell carcinoma cells, CRH production re- leads to activation of PKC-dependent pathways. There are
sponded to forskolin, an agent that raises intracellular data suggesting that CRH signal transduction is directly
cAMP levels (361). Conversely, dexamethasone inhibited coupled to calcium channels (109, 198, 388). The CRH signal
CRH peptide production in the same cell systems (361). is transduced more efficiently into stimulation of cAMP
However, despite noticeable changes in CRH peptide pro- production through CRH-R1 than CRH-R2 (74, 442). For
duction, the level of the corresponding CRH mRNA re- example, CRH and urocortin are more potent than sauvag-
mained unchanged under each one of the tested condi- ine and urotensin I in activation of adenylate cyclase activity
tions, suggesting posttranscriptional regulation. through CRH-R1, although urocortin, sauvagine, and uroten-
sin I are more potent than CRH in the stimulation of cAMP
V. CORTICOTROPIN RELEASING HORMONE production through CRH-R2 (74, 442). The affinity of CRH
RECEPTORS for CRH-R1 is similar to that of urocortin and sauvagine,
whereas CRH-R2 affinity for CRH is significantly lower
A. CRH Receptors: Structure and Function than for urocortin, sauvagine, and urotensin I (5, 74, 388).
CRH-R have so far been identified in adrenal glands,
So far, two distinct genes encoding the CRH receptor testes, ovaries, prostate, kidney, liver, gut, spleen, circu-
(CRH-R) family (CRH-R1 and CRH-R2) have been cloned lating immune cells, synovium, heart, skeletal muscle,

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 985

uterine myometrium, vascular endothelium, arterial BP) that may inactivate or inhibit local CRH signaling in
smooth muscle, endometrium, placenta, and skin (5, 13, placenta, amnion, chorion, and maternal decidua (293,
19, 74, 86, 100, 139, 150, 151, 163, 181, 190, 200, 267, 268, 297–301). Near term there is a rapid surge in circulating
296, 297, 301, 302, 305, 318, 320, 321, 334, 338, 357, 388, CRH in human pregnancy that coincides with a decrease
391, 409, 410, 456). Molecular and pharmacological char- of plasma CRH-BP levels (72, 235, 293).
acterization of CRH receptors have shown that CRH-R2 is
predominantly expressed in peripheral tissues, e.g., heart,
B. CRH Receptors: Skin
skeletal muscle, smooth muscle of arterial wall, vascular
endothelium, uterus, placenta, and immune cells (74, 139,
1. Rodent
151, 163, 190, 200, 297, 301, 318, 334, 388, 391). Notwith-
standing, CRH-R1 has been also detected in intrauterine Searching for mRNA coding for CRH receptors using
tissue, placenta, and immune cells (94, 139, 181, 297, 301, RT-PCR amplification of RNA isolated from C57BL6
318, 396, 409). The signal transduction pathway activated mouse skin has clearly shown the predicted 407-kb prod-
through peripheral CRH receptors is coupled to the stim- uct representative of the CRH-R1 mRNA (360). The tran-
ulation of cAMP production and subsequent activation of script was present throughout the entire hair cycle, even
PKA (74). However, involvement of calcium-activated during the telogen phase (360). In the hairless mice
pathways by phospholipase C or membrane-bound cal- CRH-R1 mRNA was also detected; RT-PCR assay with
cium channels has also been reported (19, 109, 186 –188, Southern blot hybridization to murine CRH-R1 cDNA
198, 242, 357, 388). In addition, the NO/cGMP-dependent showed its presence in the epidermis (357). Studies with
pathway appears to mediate CRH-R activated vasodila- full thickness skin in the C57BL6 mice showed a 2.7-kb-
tion in human fetal-placental circulation (86), and a sus- long CRH-R1 mRNA transcript by Northern hybridization;
tained vasodilator effect of CRH and sauvagine on rat it was readily detectable in anagen skin and below detect-
mesenteric artery through stimulation of NO production ability in telogen skin (360). In the same model, CRH-R1
has been reported (19). protein was localized by indirect immunofluorescence to
CRH-R located in heart and vasculature mediate the keratinocytes of the ORS, hair matrix, dermal papilla of
inotropic, vasodilatory, and antiedema effects of CRH or anagen VI hair follicles, keratinocytes of inner root sheath
CRH-related peptides (19, 74, 112, 268, 316, 320, 440, 441), (IRS), and ORS of early catagen; in contrast, skin immu-
whereas receptors present on immune cells mediate com- noreactive CRH-R1 was very low in telogen skin (Fig. 4)
plex immune/inflammatory responses (27, 305, 418, 430). (321). These results suggest that both transcription and
Although central CRH (neuroendocrine) effects are im- translation of the CRH-R1 gene are hair cycle dependent.
munoinhibitory (27, 268, 305, 418), in peripheral tissues Most recently, RT-PCR assays using primers described in
CRH effect on immune balance is complex (27, 142, 180, Reference 318 identified a 615-bp fragment common to
268a, 305, 331, 333–335, 418, 430, 440, 441). In fact, CRH the ␣- and ␤-variants of the mouse CRH-R2 cDNA (318;
may have an opposite immunoactivating effect in periph- Fig. 5). Sequencing of the cloned segment showed an
ery, by stimulating proinflammatory reactions (180, 305, almost perfect match with the published mouse CRH-R2
418, 430), although an anti-inflammatory effect for exog- genebank sequence; the single difference was a change of
enously applied CRH has also been described (142). Paez G to A found at position 675 in the known sequence (base
Pereda et al. (268a) showed that CRH can either stimulate 7 in the experimental sequence). This base permutation
or inhibit IL-1 production depending on monocyte activa- did not affect the amino acid sequence in the encoded
tion status; therefore, it is likely that activation-dependent peptide fragment that was exactly the same as the known
signal transduction through CRH receptors may explain sequence.
sometimes contradictory phenotypic effects in immune CRH did exhibit specific binding to skin, as detected
cells. Locally produced CRH may regulate steroid produc- by autoradiography with a 125I-[Tyr]-ovine CRH (oCRH)
tion in gonads and adrenal glands via a paracrine mech- tracer (321). CRH binding was localized to keratino-
anism of action (56, 102, 103, 150, 387). In placenta and cytes of hair follicles and the epidermis and dermal pan-
uterus, CRH receptors that include CRH-R1 and CRH-R2 niculus carnosus (mostly to dermal muscles; Ref. 321).
are involved in maintenance of pregnancy as well as Displacement curve analysis revealed a single high-affin-
initiation and maintenance of labor (235, 297, 301, 456). In ity binding site with a dissociation constant (Kd) of 3.62
this setting, CRH receptors are activated through para- nM and maximum binding (Bmax) of 2.42 amol/mm2 of
crine or autocrine mechanisms by locally produced CRH scanned skin section. Control determinations in rat pitu-
or urocortin (297, 301). This action requires precise time- itary revealed a single high-affinity binding site with a
and tissue-restricted differential expression of CRH re- similar Kd of 1.53 nM, but with larger number of binding
ceptors in intrauterine tissues (86, 87, 94, 297–303, 318, sites, e.g., Bmax was 18.1 amol/mm2 (321). During the hair
319, 396, 456). Regulation of CRH activity may be also cycle, 125I-[Tyr]-oCRH specific binding to isolated cell
achieved by the production of CRH binding protein (CRH- membranes was at the same level in telogen, anagen III,

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


986 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

tors. Contributing factors in determining the effect of


exogenous CRH could include the endogenous produc-
tion of CRH and CRH-related molecules and the CRH-
activated production of ACTH and MSH.
Preliminary in vivo experiments in mice showed that
CRH released slowly from subcutaneous implants placed
in the direct vicinity of telogen hair follicles effectively
resulted in follicular arrest with extended telogen (rest-
ing) stage (R. Paus, K. Fechner, and L. Mecklenburg,
unpublished data). It is not yet clear whether hair follicle
cycle arrest is a direct action of CRH or whether it simply
reflects a systemic effect of serum corticosterone. High
serum levels of CRH could substantially raise corticoste-
rone levels through stimulation of the HPA axis (444).
High systemic levels of glucocorticosteroid would clearly
suffice as the explanation for the profound anagen-inhib-
itory effect of CRH implants (cf. Refs. 272–274).
FIG. 5. Expression of the CRH receptor type 2 (CRH-R2) gene in the
B) CRH EFFECTS ON THE VASCULAR AND IMMUNE SYSTEMS.
skin of C57BL6 mouse. Primers sequence and RT-PCR conditions for
amplification of the 615-bp fragment common to the ␣- and ␤-variants of Intradermally injected CRH induces local mast cell de-
the mouse CRH-R2 cDNA were performed as described in Reference granulation in rats and mice (409). This effect is dose
318. This fragment was sequenced commercially using the BigDye chem- dependent, mediated by the amidated form of CRH, and
istry kit and a 377XL DNA sequencer from Perkin Elmer/Applied Bio-
systems (Commonwealth Biotechnologies, Richmond, VA). Two internal inhibited by the nonpeptide CRH-R1 antagonist anta-
primers, inverses of each other and with sequences GTCAATCCTGGC- larmin. Because CRH-R1 mRNA has been detected in
GAGGACGA and TCGTCCTCGCCAGGATTGAC, were used for se- mast cells, those effects may be mediated through
quence determination. The cloned segment differed from the published
mouse CRH-R2 genebank sequence only in a single base (G to A change CRH-R1 (409). CRH-induced degranulation of mast cells is
at position 675 in the known sequence; base 7 in the experimental accompanied by increased vascular permeability (re-
sequence). Lane A, DNA ladder; lane B, anagen III; lane C, anagen IV; duced by treatment with H1-receptor antagonists), leading
lane D, anagen VI.
to the suggestion that this cutaneous effect of CRH may
be primarily addressed at enhancing local vascular per-
and anagen VI skin (321, 357). It is possible that autora- meability (409). However, others have found that direct
diography may have detected both CRH-R1 and CRH-R2 topical application of CRH to cutaneous or mucosal tissue
because of the low specificity of the technique. Because had instead vasoconstrictive and anti-inflammatory ef-
the distribution of CRH-R1 is hair cycle dependent, fects (134, 237, 440). Particularly interesting are the stud-
CRH-R1 may predominate in epidermal and follicular ke- ies of CRH on animal models of tissue injury (134, 237,
ratinocytes, whereas CRH-R2 gene expression would pre- 333–335, 440, 441). Thus when CRH is injected subcuta-
dominate in extrafollicular compartments such as pannic- neously or intravenously to rats with thermal injury, fluid
ulus carnosus. accumulation is decreased by ⬎50% in injured skin of
A) CRH EFFECT ON HAIR GROWTH AND EPIDERMIS. CRH par- treated animals, independent of the functional activity of
ticipation in the regulation of hair growth was evaluated the HPA axis. This action has been explained by CRH-
in a well-characterized skin organ culture system (272, induced decrease in negative interstitial fluid pressure on
273), in which DNA synthesis in epidermal and dermal traumatized tissues, thereby reducing edema. Moreover,
compartments reflects predominantly proliferation of ker- local injection of CRH reduced doxorubicin-induced in-
atinocytes, either epidermal or follicular (357). In this flammation in the eyelid of rabbits and decreased the
model, addition of CRH to telogen and anagen IV skin was severity of skin injury (237). In the latter study, CRH did
found to stimulate DNA synthesis in epidermal keratino- not alter vascular permeability but reduced the expected
cytes, without measurable effect on the dermal compart- acute influx of monocytes and macrophages and pro-
ment (357). However, in anagen II, CRH had the opposite tected the skin overlying the injection site, substantially
activity on epidermal DNA synthesis, exerting an inhibi- reducing the extent of injury. Other studies have shown
tory effect, whereas dermal DNA synthesis was enhanced that in the rat urocortin has significantly higher potency
(357). These experiments suggest strong hair cycle re- than CRH, inhibiting heat-induced cutaneous edema
striction in the expression of CRH actions on epidermal (422). Because ␣-helical CRH-(9O49) reversed the inhibi-
and follicular keratinocyte proliferation. Thus experi- tion of edema produced by either urocortin or CRH, at
ments with exogenous CRH show variable effects de- doses that did not affect ACTH secretion, it has been
pending on the cellular population targeted and on the suggested that those effects are mediated through
hair cycle-dependent expression of CRH-related recep- CRH-R2 (422). Additional CRH effects on inflammation

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 987

include antinociceptive actions and acceleration of cytes (359, 361); the expression was upregulated by UVB
wound healing (333–335, 440, 441). irradiation or 12-O-tetradecanoylphorbol 13-acetate
There is indirect evidence for CRH activity on the (TPA) treatment (359). CRH-R1 gene expression was high
local regulation of blood flow through effects on the in melanoma and squamous cell carcinoma cells (121,
cutaneous vasculature. These would be mediated by in- 361).
teraction with CRH receptors present on smooth muscle A) BIOCHEMICAL CHARACTERIZATION OF CRH RECEPTORS. Hu-
of the arterial wall. For example, vasorelaxing effects man melanoma cells and HaCaT keratinocytes exhibit
have been shown in rat mesenteric arteries that imply specific binding sites for CRH; application of CRH or of
mediation via CRH-R2 on vascular smooth muscle (320). the related sauvagine or urocortin peptides induces rapid
Other studies have demonstrated the presence of CRH and significant increases in intracellular calcium (109,
binding sites of high and low affinity on endothelial cells: 357). The effect appears to be specific for these peptides,
Kd of 2 ⫾ 0.2 ⫻ 10⫺10 and 1.77 ⫾ 0.14 ⫻ 10⫺6 M and Bmax since the neuropeptides ␤-endorphin, ␣-MSH, and ACTH,
of 0.79 ⫾ 0.095 and 0.97 ⫾ 0.12 fmol/mg protein, respec- which also interact with G protein-linked receptors, had
tively, for high- and low-affinity binding sites (116). The either no or minimal effects (357). The calcium stimula-
expression of those receptors was accompanied by an tory activity in human melanoma cells is higher for CRH
inhibitory effect of CRH on IL-1␣-induced prostacyclin than urocortin or sauvagine peptides, being detected al-
and prostaglandin synthesis through inhibition of phos- ready at concentrations as low as 10⫺12 and 10⫺10 M in a
pholipase A2 and cyclooxygenase (114 –116). Most re- dose-dependent manner (109). Similar to hamster mela-
cently, a dual vasodilatory effect of CRH and sauvagine on noma, the CRH induced intracellular calcium accumula-
mesenteric artery was described: a short direct effect on tion in these human cell lines was inhibited by the CRH
smooth muscle, followed by sustained endothelium de- antagonist ␣-helical CRH-(9O41) and by extracellular cal-
pendent vasodilation (19). The last effect was associated cium depletion with EGTA (109). Because human mela-
with stimulation of NO production. Therefore, by analogy noma cells express CRH-R1 mRNA, it was postulated that
with the mesenteric bed (19, 114, 116, 320) and vascula- the above effects are mediated through CRH-R1 (109).
ture in the placenta and brain (86, 316, 338), it is possible Studies on the A 431 human squamous cell carcinoma
that CRH would also act through a similar pathway on cells showed that the structurally related peptides CRH,
cutaneous vascular cells, endothelial or smooth muscle, sauvagine, urotensin I, and mystixin-7 and mistyxin-11
with possible effects on local inflammation, hemostasis, stimulated cytosolic calcium accumulation and IP3 pro-
and/or coagulation. duction (186 –188). These pharmacological studies sug-
C) CRH EFFECTS ON MALIGNANT MELANOCYTES. The hamster gest therefore that the increases in cytosolic calcium are
melanoma cell line provided further insight into the mech- due to both calcium influx from the extracellular com-
anism of CRH action in the skin (109, 357). This cell partment, through calcium channels coupled to CRH re-
system expresses the CRH-R1 gene and CRH binding ceptors and pertussin toxin-sensitive G proteins, and to
sites. The melanoma CRH-R1 mRNA transcript is ⬃2.5 kb mobilization of intracellular calcium through intracellular
long, being 0.2 kb shorter than that detected in normal calcium-independent increase in IP3 (186 –188). CRH and
skin (357). CRH had demonstrable biological effects that CRH-related peptides can also stimulate cAMP produc-
consisted of a rapid and dose-dependent increase in in- tion and tyrosine phosphorylation in A 431 carcinoma
tracellular calcium concentration. This effect was re- cells, albeit with a different pattern for each group of
duced by preincubation with the CRH antagonist ␣-helical peptides tested (186 –188), suggestive of differential CRH
CRH-(9O41) and actually inhibited by depletion of extra- receptor subtype activation (187). Most recently, we have
cellular calcium with 3 mM EGTA. Therefore, CRH signal provided evidence for the existence of CRH receptors in
transduction appears to be coupled, at least partly, to human keratinocytes and defined cAMP-mediated CRH-
activation of calcium channels (109). The CRH-related stimulated pathway in these cells, with demonstrable in-
peptides sauvagine and urocortin also induced increases hibitory activity on cell proliferation (379).
in intracellular calcium concentration but at concentra- CRH receptors were also identified in human dermal
tions ⬃1,000-fold higher. fibroblasts, where two binding sites with different affini-
ties were found (115). The high-affinity binding site had
2. Human Kd of 20 ⫾ 0.22 pM and Bmax of 1.95 ⫾ 0.22 fmol/mg
protein, and the low-affinity class of receptor had Kd of
In human skin biopsy specimens, CRH-R1 mRNA was 160 ⫾ 17 nM and Bmax of 2.38 ⫾ 0.27 fmol/mg protein
identified by RT-PCR followed by Southern blot hybrid- (115). CRH receptor expression was accompanied by the
ization and found in normal scalp, compound nevus, basal phenotypic effect of CRH blockage of the IL-1␣ stimulated
cell carcinoma, and perilesional facial skin (357, 359, 361). prostacyclin (PGI2) and PGE2 production (114, 115). Be-
The CRH-R1 gene was also expressed in normal and cause CRH inhibits IL-1␣-stimulated PGI2 and PGE2 syn-
malignant melanocytes (121, 357, 359, 361) and keratino- thesis in bovine aortic endothelial cells through probable

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


988 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

inhibition of phospholipase A2 and cyclooxygenase (116), quences of the NH2-terminal fragment (NT); and exon 3
it is highly possible that a similar regulatory mechanisms codes for the COOH terminus of NT, joining peptide (JP),
may be operative in dermal endothelial cells of human ACTH, ␤-lipotropin (LPH), and for 3⬘-untranslated mRNA
skin. (17, 25, 59, 75, 99, 143, 170, 222). In the pituitary and brain,
B) CRH EFFECTS ON CELL PROLIFERATION. CRH has hetero- the primary transcripts are spliced out to generate a
geneous effects on cell proliferation. Thus, in HaCaT ke- POMC mRNA of ⬃1.1 kb (17, 25, 99, 170, 222). Shorter and
ratinocytes, cell growth is inhibited after addition of longer POMC transcripts also exist but are found only in
10⫺11–10⫺8 M CRH to the culture medium (379). In con- extrapituitary tissues that include skin (17, 76, 84, 91, 99,
trast, in melanoma cultures, the effect is biphasic in that 167, 353, 371). The POMC mRNA size heterogeneity has
short-term incubation (6 h) with CRH (10⫺10–10⫺8 M) been explained by alternative splicing, variation in the
inhibits DNA synthesis in either serum-containing or se- length of the poly(A)⫹ tail, or use of alternate transcrip-
rum-free media, whereas long-term incubation (3– 4 days) tion initiation sites (17, 76, 84, 91, 99).
at high CRH concentration (10⫺7–10⫺6 M) stimulates the POMC mRNA is translated into a single 30-kDa pro-
growth of melanoma cells (357, A. Slominski and B. tein product, the precursor of ACTH, endorphins, mela-
Zbytek, unpublished data). Both stimulation and inhibi- notropins (MSH), and lipotropins (LPH) (25, 59, 170, 383).
tion of cell proliferation by CRH have been described in These neuropeptides are differentially generated through
AT-t20 pituitary cells (239, 433), whereas cell growth is successive cell-specific processing steps and postransla-
inhibited in CRH-treated mammary cancer cells (417). tional modifications that include endo- and exopeptidase
CRH at concentrations of 10⫺10–10⫺6 M had no effect on cleavage, amidation, and acetylation (7, 8, 17, 20, 25, 59,
melanin synthesis in cultured melanoma cells (Slominski 170, 383). As indicated in Figure 6, posttranslational pro-
and Zbytek, unpublished data). cessing of POMC yields a large number of biologically
active peptides. POMC is processed in the corticotrophs
of the anterior pituitary to form ACTH, ␤-LPH, and NT,
VI. PROOPIOMELANOCORTIN
with smaller amount of ␥-LPH and ␤-endorphin (59). Fur-
ther processing occurs in the intermediate lobe: NT to
B. POMC Expression: Intracranial ␥-MSH, ACTH to ␣-MSH and corticotrophin-like interme-
diate lobe peptide (CLIP), and ␤-LPH to ␤-MSH, ␤-endor-
The pituitary gland has been recognized as an impor- phin, ␤-endorphin-(1O26) and ␤-endorphin-(1O27). All of
tant source of melanotropic factors (6, 14), defined as ␣- the peptides formed in this manner undergo amidation,
and ␤-MSH and ␤-endorphin (cf. Refs. 110, 141, 143, 201, mono- and diacetylation, phosphorylation, glycosylation,
444). The structures of the POMC gene and of the POMC and methylation before processing of POMC is completed
protein have been well established (cf. Refs. 75, 110, 141, (7, 8, 17, 20, 25, 59, 170, 383). In the pituitary, the biosyn-
143, 444). The POMC gene is expressed predominantly in thetic processing of the POMC precursor starts with spe-
the anterior pituitary where it comprises ⬃30% of all cific cleavage by the convertases PC1 and PC2 at the
mRNA (282). POMC mRNA concentration is one to two post-pairs of basic residues (Lys-Arg) and (Arg-Arg) (59,
orders of magnitude higher in pituitary than in brain (17). 90, 345, 455). The convertase PC1 cleaves POMC to pro-
In general, mammals trascribe only one POMC gene duce a 16-kDa NT, the JP, and the COOH-terminal pep-
per haploid genome (99), although the mouse genome tides ACTH and ␤-LPH with some ␤-endorphin (25, 59).
contains two nonallelic POMC ␣- and ␤-genes (425). The PC1 is expressed solely in corticotrophs determining their
mouse POMC ␣-gene is transcribed in pituitary and brain; commitment to the production and secretion of ACTH. In
it is located on chromosome 12 and shares high homology melanotrophs, the coexpression of convertases PC1 and
with human, bovine, and rat POMC genes (424). The PC2, as well as the expression of carboxypeptidase E, and
POMC ␤-gene has ⬍90% aa sequence homology (predict- the amidating and N-acetylating enzymes determine that
ed from cDNA sequence) with ␣-POMC (425). The POMC this cell type produces essentially ␣-MSH and ␤-endor-
␤-gene is not expressed in pituitary, has the characteris- phin (25, 59, 90, 106, 170, 270, 345, 383). PC2 convertase
tics of a pseudogene, and is located on chromosome 19 cleaves the first 14 aa of the ACTH sequence to generate
(425). Lower vertebrates including lamprey, fish, and ACTH-(1O14)OH peptide and ␤-LPH to ␤-MSH and ␤-en-
frogs also express two forms of POMC gene that have dorphin (25, 59). The ACTH-(1O14)OH after COOH-ter-
different degree of homology depending on the species minal amidation generates desacetyl-␣-MSH, a step re-
(12, 402). The human POMC gene is located on chromo- quired for biologic activity (25, 59, 383). Interestingly,
some 2 (25). whereas subsequent ␣-N-acetylation at the NH2 terminus,
The mammalian POMC gene contains three exons that results in ␣-MSH, enhances the melanocyte stimulat-
and two introns; exon 1 corresponds to the 5⬘-untrans- ing bioactivity of the tridecapeptide (25, 59, 383), and
lated mRNA; exon 2 contains a part of the 5⬘-untranslated acetylation of ␤-end inhibits its opioid activity (25, 59).
mRNA, sequences of the signal peptide, and start se- The O-acetylation of ␣-MSH forms N,O-diacetyl-␣-MSH. It

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 989

FIG. 6. Processing of POMC in the pituitary


gland and central nervous system. END, endor-
phin; N-TERM, NH2-terminal segment; JP, joining
peptide; LPH, lipotrophin; CLIP, corticotropin-
like intermediated lobe peptide; I, dibasic amino
acid cleavage recognition site; O, glycosylation
sites; P, phosphorylation sites. [From Castro and
Morrison (59), by permission from Critical Re-
views in Neurobiology.]

has been suggested that PC1 or PC2 may be also involved regulators of this process are dopaminergic and ␣-adren-
in the generation of ␥-MSH peptides; however, the mech- ergic agonists, PGE2, and branched amino acids (17, 25,
anism of the reaction and the potential involvement of 59, 170, 222, 329, 383).
other enzymes remain to be clarified (90, 345). The proinflammatory cytokines IL-1, IL-2, IL-6,
Production and secretion of POMC peptides is organ TNF-␣, and INF-␥ stimulate pituitary POMC gene expres-
specific, and in the pituitary it is lobe specific and under sion and production of POMC peptides (17, 25, 27, 32, 33,
multihormonal control with main stimulatory input by 118, 170, 305, 313, 383, 421). However, the mechanism of
CRH and AVP, and inhibitory regulation by adrenal glu- this regulation is complex. For example, IL-1 stimulates
cocorticoids (17, 25, 50, 59, 99, 110, 141, 170, 192, 222, 383, POMC promoter activity in AtT-20 cells in a bimodal
439). Epinephrine also has ACTH stimulatory activity, and manner (weak short-term effects followed by strong long-
the cAMP-PKA system appears to be the main intracellu- term effects) (182). A similar effect was found for TNF-␣;
lar signaling pathway for this pituitary action (10) that
IL-6 had a robust stimulatory long-term effect, whereas
stimulates ACTH secretion, similar to the effect of pitu-
INF-␥ had acute stimulatory effects followed by a marked
itary adenylate cyclase-activating peptide (PACAP) and
inhibitory effect. The effect of the cytokines on POMC
vasoactive intestinal polypeptide (VIP) (11). PACAP and
gene expression is mediated by the tyrosine phosphory-
VIP enhance separately the CRH stimulation of the POMC
gene, but without additive effect. Other factors that stim- lation cascade (182). IL-1, IL-6, and TNF-␥ can also signif-
ulate production of POMC peptides by the anterior pitu- icantly potentiate the stimulatory effect of CRH on POMC
itary are PACAP, VIP, serotonin, oxytocin, ANG II, brady- expression (17, 25, 33, 118, 170, 305, 313, 383, 421). Leu-
kinin, and leptin, whereas GABA decreases it (11, 17, 25, kemia inhibitory factor (LIF) can stimulate POMC gene
59, 81, 124, 165, 170, 222, 243, 327, 383, 414). In the expression and secretion by corticotrophs (80), acting
intermediate pituitary, neither glucocorticoids nor CRH synergistically with CRH to potentiate POMC transcrip-
has any effect on POMC gene expression, although CRH tion and production of ACTH (15). Most recently, it has
can selectively induce release of ␣-MSH (17, 25, 59, 170, been proposed that LIF and IL-1␤ activate the suppressors
222, 383). ␣-MSH release from the intermediate pituitary of cytokine signaling (SOCS) pathway, which would in-
is stimulated by ␤-adrenergic agonists, whereas negative hibit POMC gene expression and ACTH secretion, thus

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


990 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

acting as a negative-feedback mediator in cytokine activ- of POMC in lymphoid tissues, perhaps due to the pres-
ity on the neuro-immuno-endocrine axis (16). ence of distinct proteases that are selectively expressed in
LPS-treated B cells where a truncated form of ACTH is
detected (144). Neutrophils, however, have only been
B. POMC Expression: Peripheral shown to express a 9.5-kb POMC mRNA transcript and do
not release ␣-MSH (393). Recently, CRH was found to be
Transcription and translation of the POMC gene have synthesized and released by lymphocytes where it up-
been detected in most peripheral tissues including pla- regulates POMC mRNA production (144, 305, 385, 393,
centa, uterus, gonads, thyroid, pancreas, adrenals, gastro- 407). Cells in the monocytic lineage that contain POMC
intestinal tract, lungs, spleen, immune system, and skin and produce melanocortins include human peripheral
(17, 27, 32, 33, 117, 170, 220, 226, 238, 258 –260, 297, 301, blood monocytes, epidermal Langerhans cells, cell lines
305, 365, 376, 383, 445). These extracranial sites produce derived from myeloid malignancies (U937), and rodent
shorter and longer POMC transcripts, in addition to the peritoneal macrophages (55, 206, 216, 225, 238, 240, 247).
1.1-kb message (17, 84, 91, 99, 167, 353, 371). That heter- Basal production of melanocortins by monocytic cells is
ogeneity may represent alternative splicing, variation in low but increases significantly upon stimulation with LPS,
the length of the poly(A)⫹ tail, or use of alternative tran- mitogens (concanavalin A), or tumor promoters (phorbol
scription initiation sites (76). It has been questioned 12-myristate 13-acetate) (224, 247). In addition to immu-
whether mRNA smaller than 1.1 kb are translatable, since nocompetent cells, other cells known to release immuno-
they lack the signal peptide that enables the end products modulating cytokines such as cardiac muscle cells, fibro-
to exit the cell (85, 167). In this regard, a POMC mRNA of blasts, and keratinocytes have been found to synthesize
only 0.8 kb found in placenta, testes, and ovary has been and release POMC peptides and thereby contribute to the
associated with the production of secretory POMC pep- development of an immune response (220, 241, 337, 408).
tides (cf. Refs. 3, 167, 301, 383). Furthermore, in vitro In humans, increased production of POMC peptides
assays have shown that short POMC transcripts can in- has been detected during the course of arthritis, in viral
deed be translated (85); possible mechanisms allowing and parasitic infections, and in inflammatory skin dis-
translocation of peptides and proteins lacking the signal eases such as atopic eczema (32, 61, 62, 135, 136, 144, 258,
sequence have been recently described (197). It appears 259, 384, 443). ␣-MSH injected systemically has immuno-
that most tissues have the ability to produce POMC but at suppressor activity blocking IL-6- and TNF-␣-induced fe-
levels much lower than in pituitary (17, 27, 32, 33, 59, 170, ver, inhibiting the development of adjuvant-induced ar-
297, 305, 376, 383, 456). In placenta and testes, production thritis, and preventing endotoxin-induced hepatitis by
of POMC peptides is stimulated by CRH (3, 297, 301, 302). antagonizing various cytokines and chemokines (60, 61,
In the immune system, POMC expression and peptide 83, 152, 210), suggesting its participation in the immune
release are stimulated by CRH, AVP, lipopolysaccharide responses against viral infections. Indeed, POMC ele-
(LPS), and IL-1 and inhibited by glucocorticoids (27, 33, ments may bind POMC transcription factors present in
183, 305, 421). the human immunodeficiency virus-1 genome, in cyto-
Because of the significance of the immune system in megalovirus, and in some oncogenes (c-fes, MAT-1).
the skin (43), it could participate as a local site of pro- Therefore, signals known to mediate the activation of
duction of the POMC peptides ␤-endorphin, ACTH, and POMC transcription factors such as corticotropin releas-
␣-MSH. In fact, full-length POMC transcripts have been ing factor, tumor promoters, or ultraviolet light may acti-
detected in splenic mononuclear cells (MNC), which pro- vate viruses or oncogenes and thus facilitate infection and
cess POMC to ACTH through the same pathway as in possibly tumor development (208). ␣-MSH also antago-
corticotrophs (225). Also, lymphocytes and macrophages nizes the neuroendocrine effects of CRH and IL-1 (347,
form predominantly ␤-endorphin and ACTH-(1—9), with 423).
smaller amounts of ACTH-(1O24), ACTH-(1O25), ACTH-
(1O26), and ␤-endorphin-(144O146). T lymphocytes of
human and murine origin as well as lymphoma-derived C. POMC Expression: Skin
T-cell lines both express POMC mRNA and release the
POMC peptides ␣-MSH and ACTH (55, 144, 224, 225, 259, MSH was the first POMC peptide detected directly in
393). In human lymphocytes, POMC mRNA transcripts of the skin (411), which expresses the POMC gene and pro-
0.8, 1.2, and 1.5 kb have been detected (393). Similarly, duces also ACTH and ␤-endorphin (cf. Refs. 220, 365, 376,
murine ␥/␦ epidermal T lymphocytes express a truncated 378, 445). POMC gene transcription and translation in the
form of POMC and release ␣-MSH (108). The presence of mammalian skin was originally observed in the C57BL6
ACTH immunoreactivity has been clearly demonstrated in mice (370, 371). Subsequently, POMC gene expression has
rodents T-helper cells, cytotoxic T cells, and B cells (224). been found in human skin, normal and pathological, and
There is also evidence suggesting differential processing in cutaneous cell culture systems (57, 67, 71, 96, 104, 107,

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 991

TABLE1. Expression of POMC gene and production press POMC mRNA in skin, as shown by RT-PCR and
of POMC peptides in skin cells Northern blot hybridization of epidermal mRNA (G. Er-
mak and A. Slominski, unpublished data).
POMC
Cell Type mRNA POMC Peptides A) POMC GENE EXPRESSION AND POMC PEPTIDES PRODUCTION IN
C57BL6 MOUSE SKIN. The POMC transcript of C57BL6 mice
Epidermal keratinocytes Present ␣-, ␤-, ␥-MSH, ACTH, ␤-endorphin, skin is 0.9 kb long, and the POMC protein, detected with
␤-LPH
Follicular keratinocytes Present ␣-, ␤-MSH, ACTH, ␤-endorphin an anti-␤-endorphin antibody, has 30 –33 kDa molecular
Epidermal melanocytes Present ␣-, ␤-MSH,* ACTH, ␤-endorphin mass (371). The truncated form of POMC mRNA has been
Follicular melanocytes ND Absent† detected in epidermis and in epidermal Thy-1⫹ dendritic
Dermal nevocytes Present ␣-, ␤-MSH, ACTH, ␤-endorphin
Sebocytes Present ␣-, ␤-MSH, ␤-endorphin cells in the C57BL6 mouse skin (107, 108). A combination
Sweat gland cells ND ␣-, ␤-, ␥-MSH of Northern and Western blot analysis with immunocyto-
Endothelial cells Present ␣-MSH, ACTH, ␤-endorphin chemistry detected both mRNA transcript and POMC pro-
Langerhans cells Present ␣-MSH, ACTH
Monocytes Present ␣-MSH, ACTH tein during the active growth phase of the hair cycle
Macrophages Present ␣-MSH, ACTH (anagen) (122, 356, 360, 371). Because POMC gene expres-
Lymphocytes Present ␣-MSH, ACTH sion was not detected (371) or was very low (360) in the
Fibroblasts Present ND
Adipocytes ND ND resting phase of the hair cycle (telogen), the experiments
Skeletal muscle ND ACTH were interpreted as indicative of hair cycle-dependent
Smooth muscle‡ Absent Absent gene expression and thus raised the possibility of POMC
Nerves NA ␣-, ␥-MSH, ACTH
product involvement in skin appendage physiology (104,
MSH, melanocyte stimulating hormone; LPH, lipotrophin; POMC, 273, 356, 368).
proopiomelanocortin; ND, not done; NA, not applicable. * ␥-MSH is In additional studies, the exon 3 of the POMC tran-
present in transformed melanocytes. † Below detectability by immuno- script was found to be constitutively expressed in skin,
cytochemistry. ‡ Below detectability by in situ hybridization and immu-
nocytochemistry. being low in telogen and significantly upregulated in the
growth phase of the hair cycle (anagen) (360). An attempt
to determine the size of POMC mRNA on anagen skin
108, 122, 128, 131, 168, 169, 189, 211, 212, 215, 219, 223, using pituitary and brain RNA as standards give inconclu-
255, 337, 353, 354, 356, 359, 360, 363, 370, 371, 381, 408, sive results. This was due to the large discrepancy in
411, 446). Table 1 summarizes the expression of POMC tissue gene expression (per unit of mRNA ), with levels at
peptides in different cutaneous compartments. least 10,000-fold higher in pituitary than skin. Neverthe-
less, a truncated 0.9-kb POMC transcript was still present
1. Rodent throughout the entire anagen, and an additional 1.1-kb
POMC transcript was detected in anagen IV (360). An-
Immunoreactive ␣-MSH has been detected in rodent other attempt was then made to detect full-length POMC
skin from Wistar, Hooded, and Norwegian Brown rats, mRNA at a latter stage of the hair cycle (anagen VI skin)
gerbils, and hairless mice (411). Cutaneous ␣-MSH (mean (104), when melanogenic activity reaches its peak (368,
concentrations, expressed as ng/g wet tissue wt) has 369, 373). RT-PCR amplification was performed with
ranged from a low of 0.3 (Norwegian Brown rat) to a high primers spanning exons 1–3 and exons 2–3 (signal peptide
of 3.2 (gerbils), values consistently lower than the corre- and coding region) or exon 3 (coding region only), fol-
sponding hypothalamic and pituitary concentrations lowed by Southern hybridization to murine POMC cDNA.
(411). Nevertheless, the skin levels are similar to the Only a 475-bp transcript from exon 3 was detected; frag-
levels measured in extrahypothalamic-pituitary regions of ments of 810 and 738 bp corresponding to regions span-
the brain (411). Skin ␣-MSH peptide concentrations are ning exons 1–3 and exons 2–3 were present in pituitary
unaffected by hypophysectomy, at least in rat, gerbil, and and brain controls but absent in anagen skin (104). Thus,
mouse skin (411). Moreover, skin peptide concentration in anagen VI skin, the truncated 0.9-kb transcript may be
is not modified by daily injection of ␣-MSH that increase the sole or predominant POMC species (104), yet it is the
7- to 10-fold the ␣-MSH plasma levels of hypophysecto- appearance of a 1.1-kb POMC transcript in anagen IV (day
mized or intact rats. Therefore, MSH peptides detected in 5 after anagen induction) that correlates with the detec-
the skin are not of pituitary or plasma origin (411). In- tion of a 30- to 33-kDa POMC protein on same day spec-
stead, these data and direct evidence obtained in the imens; the POMC protein was not found in telogen or
C57BL6 mouse (see below) indicate that ␣-MSH present during the preceding days of anagen development (371).
in rodent skin is the result of local production. Our own Therefore, selective expression of the 1.1-kb transcript
molecular studies in mouse skin have shown that the (during anagen IV) may be the predominant pathway that
POMC gene is actually transcribed, translated (371), and leads to synthesis of the POMC protein in mouse skin
further processed to the POMC peptides ACTH, MSH, and (104). Nevertheless, truncated POMC mRNA transcripts
␤-endorphin (122, 356, 371). Also, the hairless mice ex- may be translated (85), and peptides and proteins lacking

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


992 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

the signal sequence may still be translocated (197). Thus and perifollicular nerves ending, although it is addition-
the 0.9-kb POMC mRNA cannot be entirely excluded as a ally expressed in keratinocytes of ORS and hair matrix
source of POMC peptides. during anagen IV (273). POMC mRNA has been detected
Glucocorticoids are the physiological suppressors of by in situ hybridization localized to the the dermal seba-
pituitary POMC production (99, 110, 141, 170, 383, 444), ceous glands in skin from the anagen III-VI phase of the
and correspondingly, the synthetic glucocorticoid dexa- hair cycle, whereas epithelial expression of the same
methasone attenuates POMC mRNA production in anagen POMC mRNA, first detected in scattered keratinocytes of
skin (104). Topical dexamethasone treatment is accom- the epidermis and ORS of hair follicles of anagen I, in-
panied by decreased mRNA coding for the cutaneous MC1 creases during progression to anagen VI (230, 231). The
(receptor for ␣-MSH) and for tyrosinase, and tyrosinase convertases PC1 and PC2, present at very low levels in
activity is correspondingly decreased (104). Topical ap- epidermal keratinocytes of telogen skin, increase in ana-
plication of dexamethasone also produces massive cata- gen VI, with PC1 showing the greatest change (230). In the
gen development and rapid inhibition of melanogenesis pilosebaceous unit, only PC1 is detected in telogen skin,
(104, 273, 274, 372). Thus murine hair growth and atten- whereas PC2 becomes detectable in anagen VI skin (230).
dant melanogenesis may be jointly regulated by local Thus a close temporal correlation exists between local-
POMC peptide production/MC1 receptor stimulation ization and intensity of POMC mRNA expression, expres-
(104), and suppression of this local pathway by glucocor- sion of the POMC processing convertases PC1 and PC2,
ticoids would explain the observed termination of both and detection of POMC derived peptides.
hair growth and follicular melanogenesis. B) POMC GENE EXPRESION AND POMC PEPTIDE PRODUCTION IN
In mouse skin extracts, the POMC products detected RODENT MELANOMAS AND CELL CULTURE SYSTEMS. In the amela-
by RP-HPLC coupled to specific RIA assays include ␤-en- notic rodent melanoma lines, Bomirski AbC1 hamster and
dorphin, ␣-MSH, and ACTH peptides (122, 273, 356). Thus Cloudman S91 mouse clone 6, Northern hybridization
accumulated evidence indicates that mouse skin pro- analysis detected POMC mRNA transcripts of 3.5, 1.5, and
duces POMC, which is then processed to its secretory ⬃1–1.1 kb (353). Expression of POMC mRNA in both cell
neuropeptides; triggering of this expression sequence lines correlated with detection of a 30-kDa POMC protein
may be linked to hair cycle (122, 231, 273, 356, 360, 368, by Western blotting and by cytoplasmic immunostaining
371). Accordingly, ␤-endorphin concentrations increase with antibodies against ␤-endorphin and ␥3-MSH (353,
during anagen development, reach a peak in anagen VI, 367). These results indicate that hamster and mouse mel-
and decline during follicle involution (catagen); the low- anoma cells can transcribe and translate the POMC gene.
est levels are reached in the resting phase (telogen) (122, POMC mRNA production in the AbC1 line was stimulated
371). The pattern for ACTH immunoreactivity is similar to by exposure to L-DOPA (353). Of potential oncologic rel-
␤-endorphin, e.g., low levels in telogen with a rise during evance, the amelanotic Bomirski tranplantable hamster
anagen to a peak in anagen VI skin (356). These episodic melanoma and melanotic variants are linked by common
variations in local levels of POMC peptides also correlate origin (41, 353, 367). In comparative analysis, POMC
with expression of the convertases PC1 and PC2, from a mRNA was expressed only in the rapidly growing amela-
high level in anagen VI to being low in telogen (230). The notic Ab variant, whereas POMC mRNA was not detected
pattern of predictable intermittent cutaneous accumula- in the two slower growing Ma and MI melanotic variants
tion of ACTH, MSH, and ␤-endorphin peptides indicates (353). We have therefore suggested that in this system
precise regulatory control that determines actual produc- POMC gene expression may represent an autoregulatory
tion rates; whether the peptides degradation rates are also mechanism whose expression stimulates progression to
under regulatory control remains to be tested. the malignant phenotype (367).
The differential expression of POMC products is not The melanocytic cell lines that produce ␣-MSH in-
only associated with functional changes (hair cycle clude LND2 and different subclones of B16 and Cloudman
phase) but is also anatomically restricted (cellular com- S19 melanomas, normal nontransformed (Melan A) and
partment specific). In this regard, immunocytochemistry ras transformed (LTRras11–3, PAGTori and PAGT5)
studies have shown that ␤-endorphin staining is primarily mouse melanocytes, and human melanoma lines (67, 71,
limited to sebocytes (Fig. 7), ACTH stain to epidermal and 128, 223, 457). In the B16 melanoma line, immunoreactive
ORS keratinocytes, and to skeletal muscle, whereas ␣-MSH (␣-MSH-IR) has been further identified by RP-
␤-MSH stains follicular and epidermal keratinocytes, HPLC as desacetyl-␣-MSH (223). The highest concentra-
sweat gland ducts, and sebocytes (122, 230, 231, 273, 356, tions of ␣-MSH-IR were found in the least differentiated,
371). These immunostains also show the already men- most metastatic melanoma cell lines and in ras-trans-
tioned temporal dependence on hair cycle phase, with formed melanocytes (223). In mice transplanted with var-
lowest levels in telogen and progressive increases to a ious B16 melanomas, only the highly metastatic variant
peak in anagen VI (122, 231, 371). ␣-MSH is detected was ␣-MSH positive, with the IR located predominantly in
throughout the entire hair cycle mainly in nerve bundles the peripheral invading zones (457). This was in contrast

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 993

FIG. 7. Hair cycle-associated expression of


␤-endorphin in sebaceous glands of the C57BL6
mouse. The ␤-endorphin antigen was localized
to the sebaceous glands by immunocytochemis-
try with anti-␤-endorphin antibody. Left: low
magnification (⫻100). a, Telogen; b, anagen I; c,
anagen II; d, anagen IV; e, anagen V. Right: high
magnification (⫻560) showing ␤-endorphin im-
munorectivity in individual glands. a, Anagen I;
b, anagen IV; c, anagen V; d, anagen V. [From
Slominski et al. (371), with permission from
Birkhaeuser Publishers.]

to levels below the limit of detectability of the assay in the and mouse melanoma cells and in ras-transformed mela-
low metastatic F1 variant, which grows as a noninvading nocytes (223, 353, 367, 457). Deregulated intrinsic expres-
tumor. Characterization of POMC mRNA from the S-91 sion of the POMC gene appears then as a common factor
melanoma cell line PS-1-HGPRT-1 showed three tran- preceding malignant transformation of melanocytes
scripts of ⬃6.5, 3.5, and 1.1 kb (71). These corresponded, and/or melanoma progression to less differentiated faster
respectively, to nonspliced primary transcription product, growing forms (223, 353, 367, 457). Thus local expression
alternatively spliced product, and mature POMC. In the of POMC peptides triggered by undetermined factors
same melanoma cell line, UVB stimulated POMC gene could alter melanoma cell phenotype promoting tissue
expression and production and release of ACTH and invasion. POMC peptides would exert this action through
␣-MSH in a dose-dependent manner (71). auto-, intra-, or paracrine mechanisms; alternatively,
Because all of the above models of melanoma ex- POMC peptides may affect surrounding tissues such as
press MSH receptors, the observed production of POMC immune elements or vascular system to further promote
peptides could represent an autostimulatory mechanism, tumor growth and metastatic cascade.
which accelerates growth and aggressive malignant be- In addition to being present in melanoma, POMC
havior. The general trend is that POMC gene is more peptides are also produced by several nonmelanoma ro-
actively transcribed, or produces more POMC-derived dent skin cell lines (71, 108, 220, 223). For example, the
␣-MSH in the least differentiated, most malignant hamster immortalized (BALB/c) PAM 212 keratinocytes, in which

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


994 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

production of ACTH and MSH is stimulated by UVB or 446). In this regard, cultured keratinocytes and melano-
dibutyryl cAMP (71), and immune cells such as epidermal cytes produce POMC peptides and release ␣-MSH and
Thy-1⫹ lymphocytes, Langerhans cells, and circulating ACTH into the medium (67, 189, 337, 354, 446). Produc-
cells of monocytic lineage express the POMC gene and tion of ␣-MSH and ACTH can be significantly upregulated
produce POMC peptides (108, 220). at the protein and mRNA levels upon treatment with
phorbol 12-myristate 13-acetate, IL-1, and UV light (67, 96,
2. Human 219 –221, 337, 341, 342, 446).
A) EXPRESSION OF POMC GENE AND PRODUCTION OF POMC
MSH was initially detected as immunoreactivity in PEPTIDES IN VIVO. Production of POMC peptides in vivo was
extracts of normal human skin and identified by RP-HPLC strongly suggested by the immunocytochemical studies
in isolated epidermis as the desacetyl-, monoacetyl-, and performed in skin biopsy specimens and cultured cells
diacetyl forms (411, 438); this established the background and was supported by the corresponding detection of
for further studies on the local expression of POMC pep- POMC mRNA. In situ hybridization experiments with nor-
tides. With the use of immunocytochemistry on frozen mal human skin showed strong POMC mRNA expression
sections of normal human skin, ␣-MSH was found in in epidermal melanocytes and keratinocytes, whereas in
epidermal melanocytes and Langerhans cells, and ACTH the dermis, POMC mRNA was detected at much lower
was found in differentiating keratinocytes (438). Patho- levels, in endothelial and perivascular cells (57). Thus
logical specimens of human skin evaluated by immuno- these studies confirmed previous work implicating epi-
cytochemistry also showed POMC-derived antigens in dermal and dermal cells as source of cutaneous POMC
dermal and epidermal compartments (381). The POMC peptides (57, 107, 168, 169, 189, 212, 219, 255, 337, 354,
peptides expressed included ACTH, ␣-MSH, and ␤-endor- 359, 381, 389, 408, 438, 446). Of note, ␥-MSH-IR was shown
phin, although a consistent pattern did not emerge (381). in nerve fibers in the basal layers of epidermis, in the
In normal corporal skin (obtained after surgery), POMC upper dermis, close to Merkel cells, in Meissner’s corpus-
peptides were detected in dermis but not epidermis. Der- cles, around ORS of the hair follicle, in nerve bundles of
mal POMC peptides were present in anagen hair follicles, the deep dermis, and to a lesser degree around sweat
where they accumulated in keratinocytes of ORS and hair glands and blood vessels (169, 212). Because POMC pep-
matrix. In human scalp, POMC peptides were seen in both tides have been detected in the circulating immune sys-
epidermal and follicular keratinocytes (381). The possible tem (27, 32, 33, 169, 219 –221, 247, 305), together with
discrepancy with the results of Wakamatsu et al. (438), cutaneous nerves, they may represent additional sources
who did detect POMC peptides in epidermis of normal of POMC peptides in the skin.
corporal skin, may be explained by the lower sensitivity Most resident skin cells express in situ ␣-MSH and
for detection in the Formalin-fixed paraffin-embeded tis- ACTH, ␤-MSH, ␤-endorphin, and ␥-MSH peptides (Table
sues (381). Direct RIA analyses in normal human skin 1), arising from different portions of the POMC protein
showed that immunoreactive MSH levels were indeed precursor. The peptide nature of the immunocytochemi-
very low or below the level of detectability (214). The cally detected antigens was confirmed with RP-HPLC
same group found high MSH levels in melanoma lesions, analyses that did demonstrate the presence of desacetyl-
which, together with our detection of ACTH, ␣-MSH, and ␣-MSH, mature ␣-MSH, and ACTH peptides in extracts of
␤-endorphin in a wide range of skin diseases suggest that human epidermis (438) and in perilesional and lesional
POMC peptide expression represents a disease-related skin from areas of basal cell carcinoma (363). Thus these
phenomenon (381). Such a possibility received strong analyses performed in biopsy specimens, in conjunction
support from observations in keloids, a rather homoge- with the cell culture studies, demonstrate that POMC
neous primary reactive skin disorder, where POMC prod- precursor protein is produced in situ, being processed
ucts were consistently detectable (10 of 11 cases) and locally to the final secretory peptides. Expression of MSH
localized to keratinocytes and mononuclear cells (381). and ACTH was accompanied by epidermal expression of
Because POMC mRNA was detectable by RT-PCR in bi- the POMC processing enzymes PC1 and PC2 convertases,
opsy specimens from normal scalp, basal cell carcinoma, with PC2 showing tendency to coexpress with ␣-MSH
and compound nevus, cutaneous POMC peptides appear (438). It therefore appears that POMC processing may be
to originate from local production (359). Together these cell type regulated similar to pituitary corticotrophs and
findings suggest that POMC peptides are produced and melanotrophs, i.e., skin cells may specialize in the pro-
stored locally in lesional (especially reactive) skin cells as duction of either ACTH or ␣-MSH as initially proposed
a component of the cutaneous response to injury. (376).
More recently, it was confirmed that human skin cells The detection of POMC peptides in skin is common
could actually produce POMC peptides either in vivo (57, to a broad spectrum of apparently unrelated conditions,
168, 169, 211, 212, 220, 255, 363, 381, 438) or in cell culture from the neoplastic melanocytic nevi, melanoma, basal
systems (67, 96, 107, 189, 337, 341, 342, 354, 359, 408, 438, cell carcinoma, and squamous cell carcinoma to the in-

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 995

flammatory keloids, psoriasis, and scarring alopecia POMC spillover from the skin is also suggested by studies
(381). This finding suggests that deregulated local produc- in severe atopic dermatitis and psoriasis (125, 135, 136,
tion of POMC may be a common reactive phenomenon in 252). These patients showed increased serum ␤-endor-
cutaneous pathology (381), a possibility supported by the phin levels; however, because of the previous immunocy-
results of studies on the progression of pigmented lesions tochemical detection of POMC peptides in psoriatic skin
(127–129, 131, 211, 214, 215, 255, 381). In semiquantitative (381), it was suggested that circulating ␤-endorphin could
study, POMC-derived antigens were below the limit of be generated, at least partly, in lesional skin (135).
detectability by immunocytochemistry in normal skin me- POMC peptide production was evaluated by indirect
lanocytes and in benign blue nevi (255). In contrast, com- immunofluorescence in skin biopsies from 16 patients
mon and dysplastic melanocytic nevi showed weak posi- with vitiligo lesions and 13 matched healthy controls
tivity for ␣-MSH, ACTH, and ␤-endorphin antigens in 14, (212). Keratinocytes from control skin and from involved
33, and 44% of the specimens, respectively, with a ten- and uninvolved skin of vitiligo patients expressed ␣-, ␤-,
dency for higher expression levels in dysplastic nevi and ␥3-MSH (212). MSH-IR was, however, higher in kera-
(255). In melanoma, POMC antigens were easily detect- tinocytes and extracellular compartments of involved viti-
able in ⬎50% of cases, and the stain was also more intense ligo skin than in control skin. The adnexal structures
and diffuse in the more aggressive forms: the nodular type (sweat glands and pilosebaceus orifices) showed no dif-
(pure vertical growth phase), the vertical growth phase of ference in the intensity of stain (212). Plasma ␣-MSH
superficial spreading type, acral lentiginous types, and levels were similar in vitiligo and control patients,
metastatic melanomas (211, 255). Direct measurement of whereas ␤-endorphin levels were higher in vitiligo, failing
␣-MSH with RIA found that concentrations in skin mela- to show the normal circadian rhythmic change (252).
nomas were variable from 0.21 to 2.32 pmol/g wt and in B) EXPRESSION OF POMC GENE AND PRODUCTION OF POMC-
lymph node metastasis from 0.31 to 4.25 pmol/g wt (214). DERIVED PEPTIDES IN VITRO. Expression of the POMC gene has
The ␣-MSH-IR detected in melanomas was chromato- been noted in cells of epidermal and dermal origin that
graphically heterogeneous; only a small fraction repre- include normal melanocytes and keratinocytes (57, 66, 96,
sented authentic ␣-MSH peptide, whereas most of the 107, 189, 219, 337, 354, 359, 438, 446), Langerhans cells
immunoreactivity corresponded to more hydrophobic (247), fibroblasts (408), human dermal microvascular en-
species of larger molecular weight (128, 131). The larger dothelial cells (HDMEC) (341, 342), circulating immune
molecular weight ␣-MSH-IR species probably have pre- cells (34, 35, 40, 41, 72, 258 –260), and malignant melano-
served COOH-terminal amino acid sequence and may con- cytes and keratinocytes (67, 128, 354, 363, 411). Gene
tain the whole 1–13 ␣-MSH sequence (128, 131). expression was accompanied by detection of POMC-de-
Immunocytochemistry studies performed in 30 spec- rived ACTH, ␣-MSH, and ␤-endorphin immunoreactivities
imens from basal cell carcinoma patients showed heter- in cultured cells and culture medium. Characterization of
ogeneous distribution of immunoreactive ␣-MSH, ␤-MSH, ␣-MSH and ACTH peptides by RP-HPLC analyses of cell
and ACTH peptides in tumor cells, lesional and perile- extracts in normal and malignant melanocytes and in
sional areas, epidermal and follicular keratinocytes, der- keratinocytes showed multiple forms of POMC peptides
mal mononuclear cells, and extracellular matrix (363). (128, 131, 337, 354, 363, 411, 438), which included ACTH-
The frequency of peptide detection was similar in lesional (1O10), acetyl-ACTH-(1O10), ACTH-(1O17), ACTH-
and perilesional areas and in tumor cells. RP-HPLC stud- (1O39), desacetyl-␣-MSH, and ␣-MSH (354, 363, 411, 438).
ies combined with RIA monitoring of eluted fractions Western blot analyses have confirmed the production of
documented that ACTH and ␣-MSH of lesional and peri- ␤-LPH and ␤-endorphin in normal keratinocytes (446),
lesional areas corresponded to the authentic peptides; an whereas normal melanocytes and squamous cell carci-
additional peptide eluted at the same time as desacetyl- noma cell ␤-endorphin was detected by RP-HPLC analy-
␣-MSH. POMC mRNA was also detected by RT-PCR in ses combined with specific RIA assay (354). In melanoma
extracts of perilesional and lesional skin (363). Most in- and squamous cell carcinoma cells, ACTH, ␣-MSH, and
terestingly, the MC1-R gene encoding the receptor acti- ␤-endorphin were determined by RP-HPLC-RIA (411); one
vated by MSH and ACTH was also expressed in the same of the ACTH-IR peaks eluted at the same time as ACTH-
lesional and perilesional skin, suggesting paracrine/auto- (1O13) (354). These observations suggested that process-
crine mechanisms of action for the peptides. Perhaps ing of POMC to ␤-endorphin, ACTH, and ␣-MSH is a
POMC peptides facilitate basal cell carcinoma develop- constitutive property of skin cell type, regardless of
ment and/or progression (363). whether they are benign or malignant. Production of
Ghanem et al. (127) found that melanoma patients POMC peptides in keratinocytes and melanocytes was
have increased serum levels of ␣-MSH-IR, and these could found to be under regulatory control, being stimulated by
be positively correlated with progression of clinical stage UVB, selected cytokines, and other factors (47, 64, 67, 220,
or level of invasion. This suggests that local tissue MSH 337, 446) (see below). Also, HDMEC stimulated by UVB
could eventually spill into the blood. The possibility of and UVA produce and release ␣-MSH and ACTH (341,

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


996 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

342). Of further interest, these endothelial cells express TABLE 2. Modulators of POMC gene expression
the convertase PC1, whose mRNA production is again in the skin
stimulated by UVA as well as IL-1 and ␣-MSH (341).
POMC Production
Agent (mRNA and/or Peptides)
3. Regulation
Cytokines and growth factors
IL-1 Stimulated
It is generally accepted that translation of the POMC TNF-␣ Stimulated
mRNA requires the full-length transcript composed of Endothelin-1 Stimulated
three exons, where exon 1 contains the signal peptide (17, TGF-␤ Inhibited
Physicochemical factors
99, 170, 383). In human skin cells, the size of the POMC Ultraviolet radiation Stimulated
transcript is of 1.2 or 1.3 kb, similar to that detected in the Dibutyryl cAMP Stimulated
hypothalamus and pituitary (220, 337, 446). In contrast, in TPA and PMA Stimulated
L-Dopa Stimulated
rodent skin, a shorter 0.9-kb form predominates, although Dexamethasone Inhibited
a full-length 1.1-kb transcript is also detected (104, 360,
371). It is not yet clear which of these forms is translated TPA, 12-O-tetradecanoylphorbol 13-acetate; PMA, phorbol 12-my-
ristate 13-acetate; IL-1, interleukin-1; TNF-␣, tumor necrosis factor-␣;
in rodent skin, since it is possible that the shorter 0.9-kb TGF-␤, transforming growth factor-␤.
POMC could be translatable (85) because of mechanisms
that would allow the translocation of peptides and pro-
teins lacking the signal sequence (197). Cloning of POMC central level, could play a more dominant regulatory role
mRNA fragments from human skin has shown only 86% on POMC peptides production in the skin. In this regard,
homology with the pituitary transcript, with no differ- the proinflammatory cytokine IL-1 has a significant stim-
ences in the region coding for ACTH and ␤-endorphin ulatory/inductory effect on POMC gene expression in res-
(57). This only partial POMC mRNA homology raises the ident skin cells and circulating immune cells including
question of whether additional nonallelic forms of POMC macrophages (27, 32, 33, 57, 219 –221, 337, 445, 446).
are expressed in the skin. It must be noted that lower Indeed, IL-1 increases production of POMC mRNA and
vertebrates express two forms of POMC gene (12, 402), ACTH, ␣-MSH, and ␤-LPH peptides in cultured normal
whereas in the mouse genome, two nonallelic ␣- and and malignant epidermal melanocytes and keratinocytes
␤-POMC genes have been identified (425). (67, 219 –221, 337, 445, 446). In HDMEC, expression of
Production of the final ACTH, MSH, and endorphin POMC gene is also stimulated by IL-1 (341, 342). Another
peptides is a complex multistep process that requires cytokine, TNF-␣, stimulates production of POMC mRNA
POMC processing by prohormone convertases (PC1 and in normal dermal fibroblasts (408), whereas transforming
PC2). The PC1 gene is expressed in human keratinocytes, growth factor (TGF)-␤ inhibits it in the same cell system
where it is upregulated by UVB (47, 48), whereas Scholzen and in keratinocytes, but not in keloidal fibroblasts (408).
et al. (341) showed that dermal microvascular cells ex- Neither TNF-␣ nor endothelin-1 affects POMC peptide
press PC1 mRNA, which is again stimulated by UV, and by production in melanocytes and keratinocytes (67). Fur-
IL-1 and ␣-MSH. Immunocytochemistry studies showed ther confirming the selectivity of cytokine action, the
that the PC1 and PC2 enzymes are expresssed in the release of ␣-MSH by transformed human keratinocyte line
human skin, in epidermal keratinocytes and melanocytes is not altered by the cytokines IL-2, IL-6, TGF-␤, and
(438), and in the anal mucosa (159). In the C57BL/6 interferon (IFN)-␥. Thus selected cytokines and growth
mouse, the enzymes are expressed in follicular and epi- factors produced in the skin can act locally, stimulating or
dermal keratinocytes and sebaceous gland (230). The inhibiting POMC gene expression and peptide production
overall processing of POMC in skin cells yields multiple in a cell type-restricted fashion.
forms of ACTH, desacetyl-␣-MSH, ␣-MSH, ␤-LPH, and In rodent skin, the expression of POMC gene and
␤-endorphin (122, 131, 223, 337, 354, 356, 363, 411, 438, production of POMC peptides appear to be regulated by
446), but whether similar posttranslational modifications the same biological clock that controls hair cycle (104,
of intermediates of POMC, as in pituitary and brain, e.g., 122, 273, 356, 360, 368, 371). Although the effector mes-
amidation, acetylation, glycosylation, and phosphoryla- sengers of that clock remain the subject of numerous
tion, would also occur in the skin remains to be defined. hypotheses, a pacemaker, local or distant, would generate
A list of the factors that can stimulate or inhibit the signals that periodically stimulate and inhibit cutane-
production of POMC peptides in the skin is presented in ous POMC gene expression (104, 272, 273, 276, 356, 360,
Table 2. The lack of cutaneous access to the rich source 368, 371, 372, 376). Because the anagen-induced cutane-
of neuroendocrine factors represented by the hypothala- ous POMC gene expression is inhibited by topical appli-
mus suggests that control of full expression of the POMC cation of the synthetic glucocorticoid dexamethasone,
gene in the skin is probably different from the pituitary. It which also terminates hair growth, it is possible that the
is possible that the immune system, also involved at the dexamethasone-induced attenuation of POMC gene activ-

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 997

ity may be mediated by a direct effect on the cutaneous oxidative stress, cAMP, and PKC all appear to participate
clock (104). in the regulation of POMC gene expression.
A major cutaneous stressor, UVR, can stimulate/in-
duce cutaneous POMC gene expression as shown in hu-
VII. SKIN AS A TARGET FOR
man and rodent normal and malignant cultured keratino-
PROOPIOMELANOCORTIN PEPTIDES
cytes and melanocytes (48, 64, 67, 71, 96, 220, 221, 337,
446). The response to UVR was found to be dose depen-
A role for POMC peptides in the regulation of skin
dent when assessing ACTH and ␣-MSH production and
pigmentary system is suggested by observations in the
secretion in normal and malignant keratinocytes (67, 71,
POMC knock-out mice model; surprisingly, these animals
219 –221, 337, 446) and of ␤-LPH and ␤-endorphin in ker-
survive to adulthood (450). Lack of POMC gene expres-
atinocytes (446). These effects were accompanied by cor-
sion is characterized by the main phenotypic traits of
respondingly increased POMC mRNA production . A sim-
severe obesity and adrenal insufficiency; also prominent
ilar UVR-induced stimulation of POMC gene expression,
is a defect in fur pigmentation (450). This pattern is
with increased ACTH and ␣-MSH production and release
strikingly similar to the clinical picture of patients with
to the culture media, was observed in rodent transformed
pituitary POMC gene mutations that generate allelic forms
keratinocytes and melanoma cells (64, 71). Most recently,
with defective production of POMC protein (154, 196).
UVB and UVA treatment of HDMEC resulted in increased
Affected individuals present with red hair pigmentation,
expression of the POMC gene with increased production
severe early-onset obesity, and adrenal insufficiency.
and release of ACTH and ␣-MSH (341, 342). Humans and
There is also a large body of clinical information on
horses exposed to sunlight exhibit increases in the circu-
POMC peptides excess syndromes that confirm the skin
lating levels of ␣-MSH and ACTH (129, 157, 158), and
as a target for POMC-derived peptides (cf. Refs. 101, 113,
experimental whole body exposure to UVR increases
137, 141, 171, 201, 202, 204, 220, 221, 244, 282, 283, 352,
␤-LPH and ␤-endorphin serum levels (23, 205). UV-ab-
376, 384, 397, 412, 445). For example, humans with patho-
sorbing topical sun blockers abrogate the ␤-LPH response
logically increased levels of plasma ACTH (Addison dis-
(23), implicating mediation by a photoreaction. Although
ease) or excessive ACTH production by tumors (Nelson
the pituitary gland was postulated as the source of these
syndrome) have hyperpigmentation and skin atrophy
humoral responses, the experiments cited suggest a
(101, 113, 137, 141), while the administration of MSH or
stress-sensing role for the skin and also support the skin
ACTH peptides stimulates melanogenesis (201, 202, 204).
as an alternate source for circulating neuropeptides.
Prolonged administration of synthetic ACTH in humans
Because of the significance of its interaction with the
induces acne, skin atrophy, hyperpigmentation, and hy-
skin, the contribution of UVR to the production of nuclear
pertrichosis (101, 113, 137, 141), whereas elevated serum
signals deserves further examination. Theoretically, UVR
concentrations of ␣-MSH are associated with skin pig-
could activate the genome directly or, indirectly via action
mentation (289). Additional research performed on hu-
on membrane-bound signaling processes, production of
man and animal models (see sect. VII, A and C) indicates
second messengers, or via pathways activated by oxida-
that also immune, epidermal, adnexal, vascular, and der-
tive stress (cf. Ref. 378). In support of the latter possibil-
mal structures represent additional targets for POMC pep-
ity, N-acetyl-cysteine (NAC), a precursor to glutathione
tides.
that acts as an intracellular free radical scavenger, abol-
ished the UVR-stimulated POMC peptide production in
human keratinocytes and melanocytes (67). It was then A. Melanocortin Receptors
postulated that NAC may inhibit POMC peptides produc-
tion through attenuation of the oxidative stress triggered It is known that the phenotypic effects of POMC
by UVR (64, 67). The tyrosine phosphorylation inhibitors peptides are mediated via interaction with cell surface
tyrphostin and genistein had no effect on basal or UVR- receptors linked to the G protein (101, 110, 141, 257, 444).
stimulated POMC production (67). Exposure to dibutyryl MSH and ACTH activate melanocortin (MC) receptors,
cAMP had stimulatory effects on basal ␣-MSH and ACTH whereas ␤-endorphin acts predominantly on ␮- and ␦-opi-
production in a number of cell lines, which include trans- oid receptors. The expression of MC and endorphin re-
formed (BALB/c) PAM 212 keratinocytes and Cloudman ceptors in different skin compartments is listed in Table 3;
S91 melanoma cells, PS-1-HGPRT-1 clone, and normal the following section discusses the skin MC receptors.
and transformed human keratinocytes and melanocytes, Investigations performed in frog melanophores
suggesting involvement of cAMP-activated pathways in showed that ␣-MSH interacts with cell surface receptors
that process (67, 71). Finally, phorbol 12-myristate 13- to induce cAMP production and the final phenotypic ef-
acetate and TPA enhanced POMC gene expression in fect of skin darkening (2, 141, 201, 350). In S91 melanoma
melanocytes and keratinocytes (219, 220, 337, 446), sug- cells, MSH interacts with specific cell surface receptors,
gesting involvement of PKC-activated pathway(s). Thus activates adenylyl cyclase activity, and increases intracel-

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


998 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

TABLE 3. Expression of MC and opioid receptors to a tryptophan residue (NSP-ACTH). NPS-ACTH was
found to rapidly stimulate intracellular calcium mobiliza-
Cell Type MC Receptor Opioid Receptor
tion and production of 15-lipoxygenase metabolites of
Epidermal keratinocytes MC1 ␮-Opiate receptor arachidonic acid in bovine fasciculata-reticularis cells
Follicular keratinocytes MC1 ␮-Opiate receptor (451). Therefore, calcium and lipoxygenase metabolites
Epidermal melanocytes MC1, MC2† ND may act as additional second messengers of ACTH in
Follicular melanocytes MC1 ND
Dermal nevocytes MC1 ND bovine adrenal cells, and the ACTH-(11O24) region may
Sebocytes MC5 ␮-Opiate receptor be important in this regulation (451). However, it is still
Sweat gland cells MC1, MC5 ␮-Opiate receptor unclear whether the reported effects reflect the activation
Endothelial cells MC1 Absent
Langerhans cells MC1 ND of solely MC2-R or, of other receptors as for example,
Monocytes MC1 ND MC5 that has high affinity for ACTH and is also expressed
Macrophages MC1 ND in the adrenal cortex (248, 431). Pharmacological studies
Lymphocytes MC1 ND
Fibroblasts MC1 Absent with cloned MC receptors have shown that their signal
Adipocytes MC2, MC5 Absent transduction is coupled to activation of adenylyl cyclase
(82, 88, 123, 248 –251, 256, 257). Although modest activa-
MC, melanocortin; ND, not done. † Detected by RT-PCR.
tion of IP3 production by MSH binding to MC3 was re-
ported (88, 257), there is no evidence that cloned MC1 and
lular cAMP (cf. Refs. 201, 244, 279). This is followed by MC2 receptors would be coupled to other than cAMP
increases in tyrosinase activity and melanin production second messenger generating systems (88).
(447, 448), stimulation of dendrite formation (cf. Ref. The postreceptor pathways activated by ␣-MSH bind-
101), and variable effects on cell proliferation (244, 277, ing to the MC5-R in B lymphocytes involve stimulation of
279, 284, 285, 287). On the basis of these findings, it was the Janus kinase 2 (JAK2) and signal transducers and
proposed that in rodent melanocytes the phenotypic ef- activators of transcription (STAT1) pathways (52), as well
fects of MSH are mediated by cAMP through activation of as transmission through the same phosphorylation path-
PKA and phosphorylation/dephosphorylation reactions way used by cytokines and growth factors. In cells of
(92, 101, 141, 161, 193, 244, 257, 277–279, 281, 284, 285, monocytic lineage, ␣-MSH acts on MC1-R through cAMP-
367, 448). Binding of MSH to the MSH receptor with activated pathways to suppress the activation of nuclear
subsequent activation of adenylyl cyclase was found de- transcription factor NF␬B induced by inflammatory
pendent on the presence of calcium in the receptor mi- agents (227). Activation of NF␬B may be required for
lieu; calcium binding protein (CBP) may partially activate expression of cytokines and adhesion molecules (229).
the receptor even in the absence of the ligand (105). Activation of MC1-R in immune cells inhibits NO and
Activation of PKC by ␣-MSH may also be involved in neopterin production as well as prostaglandin synthesis
stimulation of melanogenesis (51, 271), and some inves- (209, 312, 390).
tigators have suggested that MSH receptor signal trans- MSH receptors have been detected and characterized
duction may be coupled to phospholipase C-activated in rodent melanoma cells (64, 68 –70, 95, 101, 161, 175,
production of IP3 and diacylglycerol with subsequent mo- 176, 244, 265, 266, 277–279, 281, 283–285, 351, 352, 358,
bilization of intracellular calcium (51). However, studies 364, 366, 434, 435), normal and malignant human melano-
on different melanoma models have generated conflicting cytes (1, 93, 101, 119, 120, 130, 161, 164, 257, 352, 401),
results, e.g., in human melanomas ␣-MSH was found in- keratinocytes (31, 35, 64 – 66, 220, 264, 363), and other
deed to stimulate IP3 production (399), whereas in ham- cutaneous cells and extracutaneous tissues including
ster melanoma, the MSH signal transduction and subse- brain, adrenal gland, gonads, and immune cells (29, 34, 45,
quent phenotypic effect were linked to cAMP without any 79, 88, 141, 147, 209, 220, 249, 251, 312, 390). In vivo
evidence of IP3 production (366). A similar discrepancy is experiments on mice showed wide distribution of MSH
found in adrenal cortex. Some authors observed that binding sites, to almost all organs (406). The work of the
␣-MSH stimulated exclusively cAMP production without pigment biology group at Yale University has been funda-
effect on IP3 (162), whereas others showed a concentra- mental in the characterization of cell-surface MSH recep-
tion-dependent effect. At low concentrations, ␣-MSH tors (38, 95, 193, 201, 244, 277–279, 284, 285, 287, 434, 435,
stimulated IP3 production, and at high concentrations, it 447– 449). In fact, the observation that after binding the
instead stimulated cAMP production (177, 178). The latter ligand, the MSH receptor is internalized and translocated
investigators further postulated that the steroidogenic ac- to the Golgi region (434) represents the first demonstra-
tion of ␣-MSH, but not ACTH, involved stimulation of PKC tion of receptor internalization after binding of a peptide
and tyrosine kinases as the result of phospholipase C hormone. Ultrastructural and biochemical assays showed
activation (179). To further search for second messengers translocation of the MSH-MSH-R complex from the cell
of MC2 receptor (MC2-R) different from cAMP, ACTH surface to melanosomes and demonstrated the presence
was modified by the attachement of o-nitrophenylsulfenyl of actual intracellular MSH binding sites (68, 101, 244, 279,

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 999

283). MSH receptor expression was cell cycle dependent tected by RT-PCR in human skin biopsy specimens of
and restricted predominantly to G2 phase (435). This G2 compound melanocytic nevus and in cultured normal and
phase coupling of increased MSH receptor expression is malignant melanocytes (362). Of note, in human melano-
associated with increased cellular responsiveness to the cytes, ACTH appears to be more potent than ␣-MSH in
ligand (40, 64, 234, 283, 449). Factors that raise intracel- stimulation of melanogenesis (161).
lular cAMP levels such as dibutyryl cAMP and cholera MC1-R has a prominent role in regulation of mamma-
toxin can also upregulate MSH receptor expression (95). lian pigmentation (88, 101, 141, 171, 248, 250, 257, 397). In
In general, factors increasing intracellular levels of cAMP, rodents, MC1-R is encoded by the extension locus (E),
such as MSH and ACTH, upregulate expression of their which has four alleles resulting from point mutations (88,
own receptors (70, 71, 249, 291, 400). Positive cooperat- 257, 317). These allelic forms differ in their responsive-
ivity of MSH receptors has been documented in mouse ness to ␣-MSH and may even constitutively activate ad-
and hamster melanoma cells as well as in immortalized enylyl cyclase without interacting with the ligand. Similar
human keratinocytes (64, 65, 234, 364). regulatory features have been described in other species
(88, 171, 257). In humans, mutations at the MC1 locus
1. Molecular characterization generate allelic forms of the receptor which differ in
of melanocortin receptors ligand specificity and most importantly in phenotypic ef-
fect concerning hair and skin pigmentation (88, 171, 257,
Melanocortin receptors were characterized in cross- 386, 397, 428). It has also been proposed that defective
linking experiments that identified a membrane-bound MC1-R function may be associated with an increased
receptor of ⬃45 kDa in a number of melanocytic lines incidence of melanoma (429).
(101). Cloning of the MC1-R and MC2-R receptors was MC1-R is also a target for other regulatory proteins.
also performed with mRNA isolated from melanocytes For example, the agouti signaling protein (ASP), which is
(82, 250). Currently, there is a family of MC receptors that expressed in skin, inhibits the binding of ␣-MSH to
includes five members MC1-R to MC5-R, that share high MC1-R, with resulting inhibition of MSH-stimulated cAMP
sequence homology (cf. Refs. 88, 171, 251). All five genes production and melanogenesis (cf. Refs. 88, 171, 251, 257,
belong to the superfamily of seven transmembrane G 397, 401). ASP also acts as antagonists to the other MC
protein-coupled receptors (82, 88, 171, 248 –251); these receptors, MC2-R to MC5-R (cf. Refs. 88, 171, 251). ASP
are 298 –372 amino acids long and have short intracellular bound to MC1-R can inhibit cAMP-mediated activation
COOH-terminal and short NH2-terminal extracellular do- and melanogenesis, independently of exogenously added
mains. MC receptors are coupled to an adenylyl cyclase, ␣-MSH (171, 257, 401). Because signal transduction in
with possible additional coupling in MC3-R to phospho- MC1-R is coupled to Gs but not Gi protein, ASP could
lipase C (cf. Refs. 82, 88, 123, 248 –251). The MC receptors antagonize auto- or intracrine effects of endogenously
have different pharmacological profiles of activation by produced MSH or ACTH. Alternatively, binding of ASP to
MSH and ACTH peptides (82, 88, 248 –251, 290, 317, 339, MC1-R could generate changes in the receptor immediate
340). In rodents, the receptor for ␣-MSH, MC1-R, has high environment that would activate other, unrelated recep-
affinity for ␣-MSH, low for ACTH and ␤-MSH, and lowest tors.
for ␥-MSH (88, 248, 250, 251, 317). In humans, MC1-R has Agouti-related protein (AGRP), which shares se-
equally high affinity for ␣-MSH and ACTH and lower quence homology with ASP, is a naturally occurring an-
affinity for ␤-MSH and ␥-MSH (82, 88, 123, 250, 251). The tagonist of MC3 to MC5 receptors that may be involved in
receptor for ACTH, MC2-R, exhibits absolute specificity the hypothalamic control of feeding behavior (452). AGRP
for ACTH (88, 250). MC3-MC5-R show equal affinity for probably acts through competitive inhibition of ␣-MSH
␣-MSH and ACTH, whereas MC3-R or ␥3-MSH receptor binding to MC3 to MC5 receptors and inhibition of ␣-MSH-
has high affinity for ␥-MSH (cf. Refs. 88, 123, 248 –251). induced cAMP production (452). Additional antagonists
MC1 receptors have been detected in skin, brain, and of MC receptors are dynorphin peptides (313).
immune system; MC2-R in adrenal gland, adipose tissue,
and skin; MC3-R in brain, placenta, gut, and thymus; 2. Regulation of MC-R expression
MC4-R in brain; and MC5-R in brain, adipose tissue, mus-
cles, esophagus, thymus, prostate, skin, and exocrine Several years ago Pawelek et al. (283) proposed that
glands that include Harderian, lacrimal, and specialized UVB-induced melanogenesis was mediated through up-
(preputial) or nonspecialized sebaceous glands (45, 79, regulation of the MSH receptor system. This was based on
82, 88, 104, 123, 141, 161, 209, 220, 248, 250, 251, 257, 317, observations of upregulated expression of functional
362, 400, 404, 431). The distribution of MC1-R, MC2-R, and MSH receptors by UVB that amplified the melanogenic
MC5-R in skin is presented in Table 3. The mRNA for effect of exogenous MSH in a dose-dependent manner
MC1-R and MC2-R have been detected in rodent skin by both in vivo and in cell culture systems (31, 38, 40, 48, 64,
Northern blot hybridization (104). MC2-R mRNA was de- 68, 71, 283, 378). UVB action appeared to involve arrest of

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


1000 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

the cell cycle at the G2 phase, when cultured rodent tides with the most significant melanogenic activity are
melanocytes express maximal MSH receptor activity and ACTH, ␣-MSH, and ␤-MSH; the activity of ␥-MSH is low,
responsiveness to MSH (64, 283). UVB stimulates expres- and that of endorphins and CLIP is undetectable (101, 141,
sion of MC1-R in normal and malignant human and rodent 161, 236, 257, 358, 400). Structurally ACTH and MSH
melanocytes and in keratinocytes (38, 64, 68, 71, 119, 120, peptides have in common the amino acid sequence -Tyr-
378). When the skin is exposed to UVR, thermal, chemical, x-Met-x-His-Phe-Arg-Trp- that contains the tetrapeptide
or biological insults, it responds with enhanced produc- His-Phe-Arg-Trp critical for melanotropic activity (141,
tion of cytokines, growth factors, ACTH, MSH, and inter- 160). The functional relevance of ␤-MSH in mammals has
mediates of melanogenesis (365, 378). In normal and ma- been debated, since some considered it an artifact of
lignant melanocytes, many cytokines such as Il-1␣, IL-1␤, ␤-LPH proteolytic degradation (343). However, Bertagna
endothelin-1, adult T-cell leukemia-derived factor/thiore- et al. (26) have demonstrated convincingly that ␤-MSH is
doxin (ADF/TRX), INF-␣, INF-␤, INF-␥, dibutyryl cAMP, a normal product of POMC processing in human nonpi-
and the hormones ␣-MSH, ␤-MSH and ACTH can upregu- tuitary tissues. ␤-MSH stimulates melanogenesis in cell
late expression of the MC-1 gene and of functional cell- culture and skin pigmentation in humans (202, 400) and
surface MSH receptors (31, 64, 67, 70, 71, 95, 119, 120, 175, rodents (101, 201, 244, 279, 306). ␥-MSH peptides have low
176, 400). IL-1 can also stimulate MC-1 receptor expres- melanogenic activity in murine and hamster malignant
sion in normal and malignant human keratinocytes (31, melanocytes (358) as well as in frog and lizard melano-
64) and HDMEC (341, 342). ␣-MSH, ␤-MSH, and ACTH as phores (101, 141, 257). Although the overall role of ␥-MSH
well as factors raising intracellular cAMP have similar peptides in the regulation of mammalian pigmentation
effects in transformed keratinocytes (31) and HDMEC may be small, selected ␥-MSH peptides could still modu-
(341, 342). Melanin precursors such as L-tyrosine and late pigmentation indirectly, by modifying the cellular
phosphorylated isomers of L-DOPA also stimulate expres- response to other melanotropins. For example, ␥2-MSH
sion of MSH receptors and melanogenic responsiveness may potentiate the melanogenic activity of ␣- or ␤-MSH,
to ␤-MSH (233, 364, 377). Retinoic acid, while stimulating whereas ␥3-MSH, acting as a partial agonist on MSH re-
MSH receptor expression, inhibits MSH-induced melano- ceptors, can largely inhibit the melanogenic activity of ␣-
genesis (69, 265). In addition, retinoic acid and vitamin E or ␤-MSH (358).
inhibit MSH-sensitive adenylate cyclase activity in mouse A) MELANOGENIC EFFECTS IN CELL CULTURES. Studies on
melanoma cells (328). TNF-␣ inhibits MC1 expression in rodent malignant and normal melanocytes have uncov-
melanocytes (119). ered the role of MSH receptors in the regulation of mela-
nocyte functions via cAMP-dependent pathways (101,
141, 155, 201, 244, 257, 278 –286, 366). Melanogenesis is a
B. Opioid Receptors highly regulated process modified by posttranslational,
translational, or transcriptional mechanisms (101, 141,
Most recently expression of ␮-opioid receptors that 244, 257, 279). In melanoma cells, both ␣- and ␤-MSH
bind with high-affinity ␤-endorphin was detected in cul- stimulate melanogenesis activating tyrosinase as well as
tured human epidermal keratinocytes (30). In human post-dopa oxidase steps (101, 141, 257, 281, 283). In ro-
skin, in situ hybridization and immunocytochemistry has dent malignant melanocytes, ␣-MSH and ␤-MSH are more
shown that the receptors are localized to keratinocytes of potent than ACTH in stimulation of melanogenesis (101,
the epidermis and ORS of the hair follicles and to periph- 244, 279). The effect of MSH on cell proliferation is vari-
eral epithelial cells of sebaceous glands and secretory able and depends on cellular genotype (101, 161, 244, 257,
portion of sweat glands (30). The expression of ␮-opiate 277–279, 284, 285, 287, 288, 352). The finding of inhibition
receptors was downregulated by both the antagonist nal- of proliferation in amelanotic cells indicates that this
trexone and the agonist ␤-endorphin (30). Presence of the effect is unrelated to production of toxic intermediates of
related ␨-opioid receptor has been documented in human melanogenesis (244, 279, 366). MSH stimulates dendrite
and mouse skin (453). Furthermore, ␤-endorphin acts as a production through a pathway independent from that reg-
local pain modulator (148). ulating melanogenesis, although cAMP mediated (257,
366, 367). Similar MSH effects, e.g.., stimulation of mela-
C. Phenotypic Effects of POMC Peptides nogenesis, dendrite formation, and stimulation of cell
proliferation, are seen in normal cultured mouse melano-
1. Pigmentary system cytes (141, 155, 257).
In cultured human melanocytes, ␣-MSH, ␤-MSH, and
The stimulatory role of melanocortins on mammalian ACTH at concentrations in the nanomolar range or lower
pigmentation has been discussed extensively (101, 126, stimulate melanogenesis, cell proliferation, dendrite for-
141, 155, 161, 171, 201, 203, 213, 244, 257, 284, 336, 352, mation, and cAMP production (1, 161, 400). In some stud-
374). There is a general agreement that the POMC pep- ies ACTH was more potent than MSH (161), whereas in

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 1001

other reports both peptides were equipotent (1, 400). the MC1-R gene and MSH receptors (35, 47, 48, 64 – 66,
␥-MSH, which stimulates cAMP production, has no signif- 220, 264, 363). Moreover, the MSH receptor protein ex-
icant effect on melanogenesis or proliferation when pressed on the cell surface of keratinocyte is similar to
added alone at concentrations equal to or lower than 100 that characterized in pigment cells (65). Experiments in
nM in human or rodent melanocytes (400). Among the HaCaT keratinocytes showed that ␣-MSH stimulates cell
effects of ␣-MSH and ACTH on melanogenesis are in- proliferation and downregulates 70-kDa heat shock pro-
creases in concentration of tyrosinase and tyrosinase- tein (HSP70) expression (264). ␣-MSH inhibition of HSP70
related proteins (1, 400, 401) and in the eumelanin-to- expression was only observed at high calcium concentra-
pheomelanin ratio (161). ␣-MSH can also stimulate tions when keratinocytes were induced to differentiate,
attachment of human melanocytes to laminin and fi- and it was accompanied by increased sensitivity to oxi-
bronectin (161) and inhibit TNF-␣-stimulated intercellular dative stress and decreased survival (264). In agreement
adhesion molecule-1 expression in normal and malignant with that finding is the enhanced MC1-R gene expression
human melanocytes (149, 246). Mutations in MC1 recep- associated with calcium-induced epidermal keratinocyte
tor that produce unresponsiveness of epidermal melano- differentiation (64). Further studies in keratinocytes
cytes to MSH have resulted in the red hair phenotype showed that ␣-MSH can modulate cytokine production
(257, 428). (47, 48, 220, 314). Specifically, ␣-MSH stimulates produc-
B) MELANOGENIC EFFECTS IN VIVO. In rodents, application of tion of the immunoinhibitory IL-10 and inhibits the IL-1-
␣-MSH stimulates melanogenesis or switches pheomelano- induced production and secretion of IL-8 and GRO␣. It
gesis to eumelanogenesis (cf. Refs. 88, 101, 126, 140, 141, has been proposed that ␣-MSH inhibition of IL-1 effects is
257, 405). In several strains of adult mice as well as in mediated by downregulation of NF␬B (219). In human
Siberian hamsters, ␣-MSH specifically stimulates follicular adult skin, epidermal keratinocytes are negative for
melanogenesis, depending on genotype and hair cycle (53, MC1-R antigen, with the exception of perilesional kera-
54, 88, 126, 140, 141, 203, 213, 257). For example, e/e mice are tinocytes of the skin involved by melanoma (35). In con-
not responsive to MSH propigmentary action (405). ␣-MSH trast, epidermal keratinocytes in fetal skin are positive for
stimulates tyrosinase activity at the transcriptional, transla- the MC1-R (35). A role for ACTH on epidermal keratino-
tional, or posttranslational level depending on the phase of cyte function is unclear, although clinically ACTH has
hair cycle (53, 54, 101, 126, 141, 257). This is consistent with been implicated in the pathogenesis of acanthosis nigri-
the hair cycle restricted expression of melanogenesis- cans (113, 137).
related genes, protein concentration, and enzymatic ac- ␥1-, ␥2-, and ␤-MSH at the high concentration of 10⫺6
tivity (368, 369, 373). ␤-MSH also stimulates tyrosinase M stimulate mouse epidermal keratinocyte proliferation
activity, as observed in skin of newborn Syrian golden in organ culture of telogen skin, but not in anagen skin
hamsters as well as black and brown mice (306), and has (370, 375). ␤-MSH inhibits epidermal keratinocyte prolif-
actual melanogenic activity in adult guinea pigs and hair- eration in anagen IV (370, 375), whereas ␣-MSH shows the
less mice (38, 39). Overall, the promelanogenic effect of same inhibitory effect on telogen skin (273). ACTH at high
MSH peptides has been documented in a number of spe- concentrations inhibits DNA synthesis in epidermal kera-
cies (88, 101, 126, 141, 257). tinocytes (356).
In humans, the systemic administration of ␣- and The finding of increased ␤-endorphin levels in sera of
␤-MSH or ACTH stimulates skin pigmentation (202). Sub- patients with atopic dermatitis and psoriasis (125, 135,
cutaneous application of ␣-MSH enhances tanning in sun- 136) has suggested a pathogenic role for the peptide in
exposed areas of the skin (204), whereas the skin and hair those conditions. Actual ␮-opiate receptors, downregu-
pigmentation of patients with Addison disease or Nelson lated by ␤-endorphin, have been detected in epidermal
syndrome is related to increased concentration of circu- and follicular keratinocytes (30). Also the ␮-opiate recep-
lating ACTH (101, 113, 137, 257). Furthermore, increased tors antagonist naloxone can relieve itching in patients
skin pigmentation has been described in a patient having with chronic liver diseases, but it is unclear whether the
increased ␣-MSH hypersecretion without adrenal failure effect is exerted at the cutaneous or systemic levels (24).
(289). In most patients with ACTH/␣-MSH excess, hyper-
pigmentation is generalized but most prominent in sun- 3. Dermis
exposed areas.
Human dermal fibroblasts have been recently recog-
2. Epidermis nized as a target for ␣-MSH (36, 191). Thus ␣-MSH stim-
ulates synthesis and activity of collagenase/matrix metal-
Since the detection of MSH binding sites in cultured loproteinase-1 in a dose-dependent manner (191). ␣-MSH
human squamous cell carcinoma and immortalized kera- alone stimulates IL-8 release; however, when added to-
tinocytes (31, 65), several groups, with a single exception gether with IL-1, it inhibits the IL-1-induced IL-8 secretion
(400), have shown that human keratinocytes express both from fibroblasts (36). These effects may be mediated by

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


1002 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

its effects on AP-1 and NF␬B function. Thus ␣-MSH di- whereas melanocortins such as ␤-MSH and ␥-MSH, which
rectly activates AP-1, but it inhibits IL-1-induced AP-1 require the expression of other MC receptors, exhibit less
activation and stimulates NF␬B, without affecting IL-1- pronounced immunomodulatory activity. It is not yet
induced DNA binding (36). These effects may be mediated clear whether MC-1R expressing immunocytes are also
by MC1-R, because this receptor is expressed on skin affected by agouti protein.
fibroblasts (36). Acting on vascular endothelial cells ␣-MSH directly
HDMEC also express functional MC1-R as detected stimulates the adherence and transmigration of inflamma-
by molecular probing for gene expression, by binding of tory cells, a prerequisite for immune and inflammatory
125
I-labeled ␣-MSH, and by the stimulation of cAMP pro- reactions. This effect could be mediated through the ex-
duction by ␣-MSH (341, 342). ␣-MSH also stimulates in a pression of adhesion molecules such as E-selectin and
dose-dependent manner the production and secretion of VCAM (see above). E-selectin is required for the rolling of
IL-8 and GRO␣, being synergistic with TNF-␣ and IL-1 cutaneous lymphocyte-associated antigen (CLA)-express-
(342). These effects appear to be dependent on cAMP ing lymphocytes and VCAM, which mediates lymphocyte
production and activation of PKA. Thus ␣-MSH may play flattening and transmigration, is the ligand for VLA-4 ex-
a crucial role on endothelial cell function, decreasing the pressed on lymphocytes (133, 156). Therefore, ␣-MSH
adherence and transmigration of inflammatory cells, a could influence the outcome of a local immune response
prerequisite for immune and inflammatory reactions. by shifting the expression of adhesion molecules. How-
There is evidence on HDMEC that ␣-MSH, which by itself ever, in ongoing inflammatory reactions characterized by
upregulates the expression of E-selectin and vascular cell the upregulation of proinflammatory cytokines, ␣-MSH
adhesion molecule (VCAM) mRNA, inhibits the LPS-me- seems to function as a downregulatory signal.
diated upregulation of these adhesion molecules (172– An important example of the immunosuppressive ca-
174, 220, 221). These findings were confirmed at the pro- pacity of ␣-MSH is demonstrated by its effect on the
tein level and the functional level by adhesion assays outcome of the contact hypersensitivity (CHS) reaction
using T lymphocytes and cultured human dermal micro- (Fig. 8). Thus administration of ␣-MSH, systemic or topi-
vascular endothelial cells (172, 174). ␣-MSH, although
capable of activating NF-␬B by itself on endothelial cells,
blocks completely the LPS- or IL-1-mediated NF␬B acti-
vation (172–174). These findings suggest that ␣-MSH mod-
ulates the activity/availability of a transcription factor
required for VCAM and E-selectin activation.

4. Immune system

The POMC peptides ␤-endorphin, ACTH, and ␣-MSH


have strong immunomodulating potential resulting in an
overall immunosuppressive effect (32, 33, 111, 209, 269,
346). The melanocortin ␣-MSH seems to have the widest
spectrum of immunomodulating capacities. Human and
murine monocytes as well as macrophages express
MC1-R, and LPS and mitogens enhance its expression on
monocytes, whereas several cytokines are without effect
(28, 312, 390). When cultured in the presence of granulo-
cyte-macrophage colony-stimulating factor, IL-4, and
monocyte conditioned medium, human peripheral blood-
derived dendritic cells express MC-1R with the highest
levels in the fully mature cells (21, 22). It must also be
noted that whereas POMC peptides affect the function of
T, B, and NK cells, it is not yet clear whether these cells
express any of the known MC receptors. MC1-R is never-
theless expressed on cells of the monocytic lineage and
FIG. 8. ␣-MSH applied topically prevents nickel-induced contact
on cells involved in the modulation of immune and in- dermatitis. On day 0, ␣-MSH cream (100 ␮M) (left) or placebo (cream
flammatory responses such as neutrophils, mast cells, only) (right) was applied in an occlusive dressing for 2 h. Afterwards, a
endothelial cells, keratinocytes, and endothelial cells (34, patch test with nickel at concentrations 0, 0.01, 0.1, 1, and 5% was
performed. The skin was inspected and photographed 2 days later.
36, 47, 63, 147, 304). The MC1-R ligands ␣-MSH and ACTH Pretreatment with ␣-MSH inhibited the erythematous response to
affect the function of cells involved in immunoregulation, nickel.

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 1003

cal, suppresses CHS at both the elicitation and sensitiza- would be in allergic diseases such as contact dermatitis
tion phases (138, 153, 209, 315, 349). In addition, when (see Fig. 8), atopic dermatitis, asthma, autoimmune dis-
␣-MSH is applied before epicutaneous sensitization with eases, and organ transplantation. In contrast, cutaneous
haptens, it induces tolerance, which is hapten specific, tumors may be associated with overproduction of POMC
since animals can be still sensitized with other antigens. peptides, which may represent a mechanism for escape
The induction of hapten-specific tolerance by ␣-MSH may from immune surveillance (127–129, 223, 353, 363, 367). In
be mediated by IL-10, because it is blocked by adminis- these cases, neutralization of ␣-MSH could represent a
tration of anti-IL-10 antibodies (138). Furthermore, in the useful approach.
mouse, a single intravenous ␣-MSH injection results in In addition to causing skin damage, UVR is associ-
significant elevation of IL-10 plasma levels lasting for ated with cutaneous and systemic immunosuppression
more than 2 wk (220). Another mechanism involved in (220, 378). Thus UVR impairs the function of epidermal
␣-MSH tolerance induction may be its downregulatory immunocompetent cells such as Langerhans cells and
effect on the expression of accessory molecules such as induces the production of immunosuppressing cytokines.
CD86 and CD40 on antigen-presenting cells (21, 22, 28). In the preceding sections we reviewed the stimulatory
Therefore, ␣-MSH may impair the function of antigen effect of UVR on the cutaneous production of POMC
presenting cells (APC) and shift the outcome of an im- peptides and on expression of MC1-R by skin cells. These
mune response toward tolerance acting directly as well as data in conjunction with expression of functional MC1-R
via IL-10 induction. The notion that ␣-MSH interferes with on immune cells implicate POMC-MC1-R axis as an im-
the function of APC has received support from experi- portant coordinator in the response to UVR. Such effect
ments with Langerhans cells prepared from normal mu- addressed probably at attenuating the UVR-induced in-
rine skin. The cells were treated in vitro with ␣-MSH flammatory reaction could be exerted directly or indi-
derivatized with dinitrofluorobenzene (DNFB) and then rectly. In the latter, UVR-stimulated ␣-MSH expression
used to induce CHS in naive syngeneic mice. Treatment would induce production of immunosuppressive factors,
with ␣-MSH impaired Langerhans cell ability to induce resulting in local (skin) and systemic immunosuppression
CHS but failed to induce tolerance, suggesting that addi- (220, 378).
tional pathways are responsible for antigen tolerance in-
duction (89); thus a second application of the allergen 5. Hair
induces production of proinflammatory cytokine IL-1
which upregulates the expression of adhesion molecules Hypertrichosis is a process whereby a minute, non-
on endothelial cells. Pretreatment with ␣-MSH before pigmented vellus hair follicle (HF) is converted into a
challenge downregulates the expression of adhesion mol- large terminal HF that produces strong, pigmented hair
ecules on endothelial cells (172). Therefore, inflammatory shafts. To some extent this process may, at least in part,
cells required for the CHS reaction are unable to adhere reflect a change in HF cycling, since hypertrichotic HF
and penetrate through the vessel wall, and CHS is sup- have a substantially longer anagen phase than the original
pressed. These findings suggest a therapeutic role for vellus HF (cf. Ref. 272). In humans, overproduction of
␣-MSH in cutaneous immune responses. ACTH, e.g., by pituitary tumors or therapeutic administra-
There is evidence from several studies indicating that tion of ACTH, is a well-recognized cause of acquired
most of the anti-inflammatory effects of ␣-MSH are ex- hypertrichosis (272, 444). This induction of hypertrichosis
erted by its NT tripeptide (209, 210). This tripeptide can by ACTH provides circumstantial evidence that the neu-
actually compete with ␣-MSH for MC-1R binding and, by ropeptide may stimulate and/or prolong anagen. ␣-MSH
itself, induces IL-10 production by monocytes as well as may also have a role in the regulation of hair growth,
downregulates CD86 expression (29). In vivo the NT MSH since ␣-MSH receptors have been detected in the human
tripeptide is capable of blocking the induction as well as HF (273).
the elicitation of CHS, similar to the intact ␣-MSH mole- In minks and weasels, the effect of POMC peptides
cule. In addition, epicutaneous application of the tripep- on hair growth has been the subject of detailed studies
tide at the site of sensitization suppresses the elicitation (cf. Refs. 126, 272). Thus it has been observed that hy-
of CHS and induces tolerance (220). Therefore, topical pophysectomy prevents the spontaneous onset of molt-
application of these tripeptides, of ⬃350 Da molecular ing, an effect corrected by exogenous ␣-MSH or ACTH
mass, and thus capable of penetrating the epidermis, may (325, 326). Furthermore, in the mink, bilateral adrenalec-
represent a novel approach for the treatment of allergic tomy causes a sharp rise in pituitary ACTH release and
and inflammatory skin diseases. premature onset of anagen (322, 324). In fact, hair growth
␣-MSH as well as its NT tripeptide can apparently in mink can be induced directly by intracutaneous injec-
serve as antagonist of proinflammatory signals and tion of ACTH, but not ␣-MSH (323) (Fig. 9). In the C57BL6
thereby downregulate immune and inflammatory re- mouse, ACTH has a bimodal hair cycle-dependent effect
sponses. Therefore, its major therapeutical potential on hair growth, e.g., intracutaneous injection of ACTH in

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


1004 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

and expression of MC1-R (104, 122, 356, 360, 370, 371);


and in hair follicle innervation (276). It has been proposed
that the hair cycle-dependent local production of POMC
and expression of the corresponding receptors are inte-
gral effectors of the biological clock that regulates cyclic
activity of hair follicles, skin, and subcutis. In this context
POMC peptides imported from extracutaneous sources
may have limited effects because of competition for local
receptors with endogenous peptides; therefore, any ef-
fect(s) of systemic supplies would be restricted to the
phases when local production is low or absent. For ex-
ample, ␥- and ␤-MSH at pharmacological doses stimulate
DNA synthesis in the epidermal compartment of mouse
skin (370, 375). However, this effect is seen only in telo-
gen but not during anagen development. Furthermore,
pharmacological doses of ␣-MSH inhibit proliferation of
epidermal and follicular keratinocytes, but again only in
telogen and not anagen skin (273). ACTH also acts in
telogen skin, where it has a dose-dependent dual effect,
e.g., at physiological doses it stimulates DNA synthesis in
dermal but not epidermal compartments, whereas at
pharmacological doses ACTH inhibits DNA synthesis in
both dermal and epidermal compartments (356).

6. Sebaceous glands

ACTH, ␣-MSH, and ␤-LPH influence sebaceous gland


function (412). In the rat, ACTH has sebotropic activity
that has been explained by its stimulation of adrenal
androgens and progesterone secretion, since adrenalec-
FIG. 9. ACTH induces anagen development in minks. Minks were
injected intradermally with ACTH (A: arrows) or with ␣-MSH (B: ar- tomy completely inhibits this sebotropic effect. Also in
rows). The contralateral hip of each mink was injected with solvent rodents ␣-MSH stimulates sebum secretion and lipogene-
control (PBS). [From Rose (323), with permission from Blackwell Sci- sis in cutaneous sebaceous glands and in specialized pher-
ence, Inc.]
omone-producing preputial glands acting as trophic fac-
tor for those glands. ␣-MSH action on the preputial glands
telogen skin induces anagen development (275), whereas of rodents stimulates production and release of female
in anagen skin it induces the premature onset of catagen sex attractant odors and of male aggression-promoting
(228). pheromones (412). ␣-MSH specifically stimulates wax and
The C57BL6 mouse has been also studied as a model sterol ester biosynthesis (79, 412). It has been postulated
for the effect of POMC peptides on hair growth (228, 273, that the sebotropic activity of ␣-MSH and ACTH is medi-
275, 356, 370, 371, 375). In mice, hair growth is synchro- ated by a peptide sequence different from the melano-
nized along three developmental stages: resting (telogen), tropic one (412). In rodents, ␣-MSH and perhaps ACTH
growth (anagen), and regression (catagen) (272, 368). The may be important for skin thermoregulation, preventing
anagen follicle produces the hair shaft formed by pig- overwetting of hairs, and by behavior regulation through
mented, keratinized epithelial cells. Hair growth and the their action on nonspecialized and specialized sebaceous
cyclic activity of the hair follicle are timed by a “biological glands (412).
clock” of unknown nature through as yet ill-understood MC5-R that are fully functional are actually ex-
tissue interactions. In rodents, the cyclic activity of hair pressed in cutaneous sebaceous and preputial glands and
follicles is accompanied by periodic changes in the phys- in Harderian and lacrimal glands (79, 431). In the latter,
iology of the entire skin (cf. Refs. 272, 368). These are melanocortin binding sites have been detected and ACTH
expressed morphologically as skin thickening, increased and ␣-MSH stimulate lacrimal protein production and
vascularization, and skin pigmentation. The cyclic activity peroxidase secretion (79, 431). ACTH also stimulates por-
of hair is coupled to cyclic changes in local immune phyrins synthesis by Harderian glands (79). MC5-R-defi-
system (272, 273); in the pigmentary system (368, 369, cient mice have severely defective cutaneous water repul-
373); in POMC expression, production of POMC peptides, sion and thermoregulation, from impaired sebaceous

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 1005

lipids production (79). In the MC5-R knock-out mice, skin cells and circulating immune cells (27, 32, 33, 67,
Harderian gland function is also impaired (79), suggesting 219 –221, 337, 341, 342, 445, 446). There is parallel stimu-
a role for MC5-R in the protection of the eye from envi- lation of PC1 gene expression (340, 341), and IL-1 also
ronmental stresses. stimulates MC1-R expression (31, 64, 119, 220, 341, 342).
In humans, it is likely that MSH and ACTH peptides Another cytokine, TNF-␣, stimulates POMC gene expres-
act on sebaceous glands functions, since hypopituitarism sion in dermal fibroblasts (408); TGF-␤ has the opposite
is associated with decreased sebum production (137). effect, inhibiting POMC gene expression in the same cell
system and in keratinocytes (408). Other related products
such as endothelin-1, ADF/TRX, INF-␣, INF-␤, and INF-␥
VIII. REGULATION OF THE CUTANEOUS can also stimulate the expression of functional MSH re-
CORTICOTROPIN RELEASING HORMONE- ceptors on melanocytes (31, 64, 70, 71, 95, 119, 120, 175,
PROOPIOMELANOCORTIN SYSTEM 176, 400). Thus, similar to UVR, the SIS also participates,
through the opposing action of selected cytokines, in the
A. UVR regulation of local expression of POMC and its corre-
sponding receptors.
UVR also has a dominant stimulatory role on CRH
and POMC peptide production as well as in the expres- C. Hair Cycle
sion of the corresponding receptors amplifying the phe-
notypic effects of MSH. Thus UVR stimulates CRH pep- In rodents, this cyclic process is characterized by
tide production in human melanocytes in a dose- dramatic changes in skin physiology, with variations in
dependent manner with increased production at higher CRH content, expression of the POMC gene, production
levels of radiation (355). UVR also stimulates POMC gene of POMC peptides, expression of CRH-R1 and MC1-R
expression with enhanced production and secretion of receptors, and expression of PC1 and PC2 convertases
POMC peptides such as ACTH, ␣-MSH, ␤-LPH, and ␤-en- (104, 122, 230, 273, 321, 356, 360, 368). Thus the local
dorphin in a number of rodent and human cutaneous cell pacemaker that determines the hair cycle also deter-
systems (48, 64, 67, 71, 96, 219 –221, 337, 341, 342, 446). mines, directly or indirectly, the expression of the CRH-
Also stimulated by UVR is the expression of prohormone POMC system in rodent skin (104, 272, 273, 276, 356, 360,
convertase PC1, responsible for the processing of both 368, 371, 372, 376). The effector messengers regulating
pro-CRH and POMC (341). Recent data suggest that UVR- cyclic follicular activity remain to be defined (104, 272,
stimulated CRH and POMC peptide production may be 273, 276, 368, 376).
mediated by UVR-triggered oxidative stress (64, 67, 378).
Thus the powerful cutaneous stressor represented by
UVR stimulates all of the components of the regulatory D. Disease States
arm (production of CRH and POMC peptides) in the local
CRH-POMC system. In addition, UVR activates the func- In normal corporal human skin (whole body except
tional arm of the system by inducing expression of the scalp and face), expression of the POMC gene is low or
corresponding receptors. The latter effect is illustrated by below the level of detectability (214, 381). In contrast,
the stimulation of MC1-R mRNA with increased expres- gene expression becomes readily detectable in patholog-
sion of functional MSH receptors that follows, in a dose- ical conditions such as keloids, psoriasis, basal and squa-
dependent manner, UVR exposure (38, 64, 68, 71, 119, 120, mous cell carcinomas, vitiligo, and melanomas (127, 131,
378). The UVR stimulatory effect on functional MSH re- 135, 211, 212, 255, 363, 381). Most interestingly, malignant
ceptors has been correlated with amplification of the progression of pigmented lesions is accompanied by in-
melanogenic effect of MSH peptides in number of cell creased expression of POMC peptides in the atypical
culture systems (31, 38 – 40, 64, 68, 71, 283); it is also seen melanocytes of dysplastic nevi or melanoma (211, 255,
in vivo, where it is associated with melanocyte prolifera- 381). Concordantly, increased production of POMC pep-
tion (38, 39, 378). tides is found in the more advanced and aggressive forms
of melanoma (127, 128, 131, 250, 353, 367, 457). Thus
disease process can stimulate cutaneous POMC gene ex-
B. Cytokines
pression.

These secretory products of the skin immune system


(SIS) also regulate cutaneous CRH-POMC system activity. E. Cellular Metabolism Mediators
The cytokine IL-1 has significant stimulatory effects on
POMC gene expression and on the production of POMC- Factors that raise intracellular cAMP levels such as
derived ACTH, ␣-MSH, and ␤-LPH peptides by resident dibutyryl cAMP, forskolin, and MSH stimulate production of

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


1006 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

CRH and of POMC-derived peptides in skin cells (64, 67, 341, With regard to the POMC peptides MSH and ACTH,
361). In addition, dibutyryl cAMP and the hormones ␣-MSH, these stimulate epidermal and follicular melanogenesis,
␤-MSH, and ACTH stimulate expression of the functional melanocyte proliferation, and modify immune and secre-
cell surface MSH receptors in melanocytes, keratinocytes, tory functions of pigment cells (cf. Refs. 64, 101, 126, 141,
and dermal microvascular endothelial cells (31, 64, 70, 71, 149, 161, 202, 220, 365, 374, 376, 445). ACTH, ␤-endorphin,
95, 119, 120, 175, 176, 341, 342, 400); in the latter, phorbol and especially ␣-MSH are strong immunomodulators ex-
12-myristate 13-acetate and TPA enhance POMC gene ex- erting an overall immunosuppressive effect (32, 33, 111,
pression (219, 220, 337, 446). Together these observations 209, 220, 269). For example, ␣-MSH can function as an
suggest that PKA and PKC mediate significant regulatory antagonist of IL-1 and induce production of immunosup-
pathways in cutaneous CRH-POMC system. pressive cytokines while suppressing the expression of
accessory molecules on APC (22, 28, 58, 138, 147, 219, 220,
254, 314). Of clinical significance, ␣-MSH and its NH2-
F. Glucocorticoids
terminal tripeptide can prevent the induction of the con-
tact hypersensitivity reaction and even block its elicita-
Dexamethasone inhibits CRH production in cultured
tion (138, 153, 209, 220, 315, 349). ␣-MSH has already been
human skin cells and attenuates POMC gene expression
used successfully for the treatment of contact dermatitis
in rodent skin (104, 361). Dexamethasone also inhibits
(Fig. 8). MSH affects epidermal keratinocyte proliferation
expression of the MC1-R gene in rodent skin (104), and it
stimulating cell differentiation and differentially modu-
terminates anagen-associated melanogenesis by inducing
lates cytokine production (cf. Refs. 64, 220, 264, 273, 314,
catagen (104, 274, 372). Thus, in analogy with their central
370, 375). ACTH stimulates hair growth by inducing ana-
action, glucocorticoids attenuate the local CRH/POMC
gen development (Fig. 9) (273, 275, 322, 323). The MC5-R
system activity.
knock-out mice model, which has defective sebaceous
gland function and thermoregulation (79), provides fur-
IX. PHYSIOLOGICAL SIGNIFICANCE OF ther evidence for the role of ACTH/MSH in those func-
CUTANEOUS PRODUCTION OF tions (79, 412, 431). ␣-MSH has a regulatory role in rodent
CORTICOTROPIN RELEASING HOMRONE nonspecialized and specialized sebaceous gland function;
AND PROOPIOMELANOCORTIN it stimulates lipogenesis and sebum secretion as well as
production and release of female sex attractant odors and
As noted above, local production of CRH and POMC male aggression-promoting pheromones (79, 412, 431).
and corresponding receptors expression can occur in re- Finally, MSH stimulates dermal fibroblast collagenase
sponse to environmental stimuli that do not reach the production and secretion as well as cytokine production
brain (e.g., UVR). This evidence strongly implies that (36, 191, 220). In summary, cutaneous POMC peptides
cutaneous production of these peptides, and their effect have marked effects on the pigmentary, immune, epider-
on skin function may be regulated independently from mal, adnexal, and dermal functions of the skin, enhancing
their systemic counterpart. The effects of the CRH/POMC melanogenesis, stimulating cellular differentiation, stimu-
responses are cell-compartment and function specific. lating hair growth and sebaceous gland functions, and
For example, CRH and related peptides stimulate intra- attenuating inflammatory reactions.
cellular calcium accumulation in malignant melanocytes
and in normal and malignant keratinocytes and affect cell
X. HYPOTHESIS ON AN ORGANIZED SYSTEM
proliferation (109, 186 –188, 357, 379). CRH action is nev-
MEDIATING SKIN RESPONSES TO STRESS
ertheless modified by a timing component, since implan-
tation of slow-releasing CRH capsules in the vicinity of HF
with the telogen phase will effectively arrest hair cycle A. Hypothesis
progression, producing an extended telogen phase (273).
Overall, peripheral CRH acts as a growth factor for the The findings in the skin of CRH/CRH-like peptides,
epidermis and HF, has antiedema effects, and is a vaso- POMC peptides, and their corresponding receptors
dilator agent potentially enhancing blood flow through strongly advocate against random occurrence (Fig. 10).
dermis acting on the vascular smooth muscle and endo- Rather, similar to the central levels, expression of these
thelium (19, 86, 114, 116, 134, 316, 320, 338, 422, 440). CRH elements is highly organized. The functional purpose of
could be important for local hemostasis, since it inhibits this CRH/POMC system would be to respond to external
IL-1-induced prostacyclin and prostaglandin synthesis and internal stresses through local pigmentary, immune,
(114, 116). It is also considered an immunomodulatory epidermal, adnexal (HF and sebaceous, eccrine, and
factor (9, 134, 142, 237, 268a, 305, 334, 335, 418, 430, 440), apoccrine glands), and vascular structures to stabilize
consistent with its ability to induce mast cell degranula- skin function and prevent disruption of internal ho-
tion (409). meostasis. This skin stress response system (SSRS) has

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 1007

FIG. 10. Diagramatic presentation of cutaneous CRH/


POMC system. A: cutaneous sites of CRH and POMC pep-
tide production. B: cutaneous sites of CRH and POMC
receptor expression. C: scheme of CRH/POMC function in
the response to injury, through para- and autocrine mech-
anisms.

the capacity to regulate its level of expression indepen- tary system; in rodents (mostly nocturnal animals), the
dent of the CNS. The SSRS operates at the cellular, tissue, SSRS is more reactive to chemical insults and coupled to
and organ level via a combination of intra-, auto-, and the activity of adnexal structures.
paracrine mechanisms (Fig. 10). Because of the differ-
ences in skin structure between furry animals and hu-
mans, as well as the nature of the predominant stress B. Background to the Organization and Function
reaching the viable portion of the skin, the operational of the SSRS
organization of the local CRH-POMC system has charac-
teristic species specificity. In humans, the SSRS is most Physiologically, the skin is the body organ with the
reactive to solar radiation and thus linked to the pigmen- most extensive exposure to the environment; as such, it

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


1008 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

has a critical role in the protection of global homeostasis. CRH/POMC molecules similar to those produced in
Thus the continuous barrier represented by skin is ex- the skin are being released continuously by the HPA axis
posed to extreme temperatures, radiation, biological and are likely to influence skin function. However, some
agents, and trauma. The most important cutaneous stress degree of cutaneous insensitivity to CRH-POMC peptides
is the solar radiation, particularly the ultraviolet spectrum of central origin may result from local impermeability or
which interacts with epidermal components and leads, in situ degradation of peptides. In contrast, an opposite
when excessive, to the local accumulation of toxic prod- endocrine central effect from the low constitutive activity
ucts. Exposure of humans to high-intensity solar radiation of the CRH-POMC skin system would seem less likely.
results in a highly circumscribed immediate reaction, i.e., Thus, when the intensity of skin damage is major, by the
typical erythematous eruption restricted to the exposure time the extent of local involvement might release a large
site; more chronically, it results in pigmentation. The amount of POMC peptides that could spill into systemic
reactive accumulation of the pigment melanin, which acts circulation, the systemic reaction to stress via the central
as a light filter and scavenger of cellular toxins and free HPA axis would have been already activated. Therefore,
radicals, prevents further skin damage. peripheral contribution to the acute systemic response to
CRH, ACTH, ␤-endorphin, and MSH peptides, to- stress would be insignificant. Nevertheless, we cannot
gether with receptors for CRH and the POMC peptides, exclude the possibility that in this setting, as well as in
are all present in the skin, while the corresponding genes pathological conditions resulting in chronic, sustained
are also expressed locally (Fig. 10). That the components CRH/POMC skin overproduction, the same molecules
of the highly organized system involved in the systemic could serve a different role as messengers of cutaneous
response to stress would be expressed in cells whose distress. The latter function might be accomplished di-
normal destiny is to differentiate and die strongly support rectly by the hormonal products themselves, or indirectly
functional relevance for this energy-requiring function. by stimulating ascending neural pathways.
Indeed, testing the functional modulation of the skin
We thank Drs. Z. Abdel-Malek, T. Kishimoto, E. Linton, J.
CRH-POMC system has shown strong linkages to UVR,
Pawelek, J. Roberts, and E. Wei for their valuable comments.
cytokines, hair cycle, disease states, and metabolic mod-
The excellent secretarial work of T. Hermann is acknowledged.
ifiers. This work was supported by National Science Foundation
Grants IBN-9896030, IBN-9604364, and IBN-9405242; American
C. Functional Organization of SSRS Cancer Society, IL Division, Grant gg-51; Bane Charitable Trust
Grant LU#9178 (to A. Slominski); Deutsche Forschungsgeimein-
When the normal integrity of the skin is disrupted, it schaft Grant Pa 345/6 –1 (to R. Paus); a grant from Wella, Darm-
is most important to restrict maximally the consequent stadt (to R. Paus); and a grant from Deutsche Forschungsgei-
biological reaction, to preserve its functional properties. meinschaft (to T. Luger).
The actions of the CRH-POMC system at the cellular level Address for reprint requests and other correspondence: A.
prevent spreading of the field of stress. Within this con- Slominski, Dept. of Pathology, Univ. of Tennessee Memphis,
Rm. 576 BMH-Main, 899 Madison Ave., Memphis, TN 38163.
text, the space-restricted paracrine, autocrine, and intra-
crine mechanisms of action are critical for limiting tissue
damage. Expression of truncated products lacking the REFERENCES
signal peptide that would allow easy export from the cell
1. ABDEL-MALEK Z, SWOPE VB, SUZUKI I, AKCALI C, HARRIGER
would also ensure that the effect would be localized to MD, BOYCE ST, URABE K, AND HEARING VJ. Mitogenic and
structures in the immediate vicinity of manufacturing melanogenic stimulation of normal human melanocytes by mela-
sites. The SSRS immunosuppressive activity decreases notropic peptides. Proc Natl Acad Sci USA 92: 1789 –1793, 1995.
2. ABE K, BUTCHER W, NICHOLSON WE, BAIRDE CE, LIDDLE RA,
recruitment of local immunocytes; local immunoglobulin AND LIDDLE GW. Adenosine 3,5-monophosphate (cyclic AMP) as
and cytokine production would abate, and local dermal the mediator of the actions of melanocyte stimulating hormone
capillary permeability would return to basal levels. Most (MSH) and norepinephrine on the frog skin. Endocrinology 84:
362–368, 1969.
important in this conception is the timing of the func- 3. ACKLAND JF, SCHWARTZ NB, MAYO KE, AND DODSON RE. Non-
tional sequence, since CRH-POMC activation must follow steroidal signals originating in the gonads. Physiol Rev 72: 731–787,
the primary reactive mechanisms. It is possible that the 1992.
4. ADLER GK, ROSEN LB, FIANDACA MJ, AND MAJZOUB JA. Protein
full expression of CRH-POMC effects may require the kinase-C activation increases the quantity and poly(A) tail length of
presence of stress-dependent products acting as potential corticotropin-releasing hormone messenger RNA in NPLC cells.
cofactors. An effect of stress-released molecules acting as Mol Endocrinol 6: 476 – 483, 1992.
5. AGUILERA G. Corticotropin releasing hormone, receptor regula-
cofactor for CRH-POMC phenotypic effects is supported tion and the stress response. Trends Endocrinol 9: 329 –336, 1998.
by, for example, the finding of pigmentation restricted to 6. ALLEN BM. The results of extirpation of the anterior lobe of the
the solar radiation-exposed areas in patients with adrenal hypophysis and of the thyroid of Rana pipiens larvae. Science 44:
755–758, 1916.
destruction (Addison disease) or with tumoral pituitary 7. ALLEN RG, HATFIELD JM, STACK J, AND RONNEKLEIV O. Post-
ACTH hypersecretion (Nelson syndrome). translational processing of proopiomelanocortin (POMC)-derived

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 1009

peptides during fetal monkey pituitary development. Adrenocorti- stimulating hormone revisited. Proc Natl Acad Sci USA 83: 9719 –
cotropin (ACTH) and ␣-melanotropins (␣-MSHs). Dev Biol 126: 9723, 1986.
164 –172, 1988. 27. BESEDOVSKY HO AND DEL REY A. Immune-neuro-endocrine in-
8. ALLEN RG, ORWELL E, KENDALL JW, HERBERT E, AND PAXTON teractions: facts and hypotheses. Endocr Rev 17: 64 –102, 1996.
H. The distribution of forms of adrenocorticotropin and beta-en- 28. BHARDWAJ RS, BECHER E, MAHNKE K, HARTMEYER M,
dorphin in normal, tumorous, and autopsy human pituitary tissue: SCHWARZ T, SCHOLZEN T, AND LUGER TA. Evidence for the
virtual absence of 13K adrenocorticotropin. J Clin Endocrinol differential expression of the functional alpha-melanocyte-stimu-
Metab 51: 376 –380, 1980. lating hormone receptor MC-1 human monocytes. J Immunol 158:
9. ANGIONI S, PETRAGLIA F, GALLINELLI A, COSSARIZZA A, 3378 –3384, 1997.
FRANCESCHI C, MUSCETTOLA M, GENAZZANI AG, SURICO N, 29. BHARDWAJ RS, SCHWARZ A, BECHER E, MAHNKE K, ARA-
AND GENAZZANI AR. Corticotropin-releasing hormone modulates GANE Y, SCHWARZ T, AND LUGER TA. Pro-opiomelanocortin-
cytokines release in cultured human peripheral blood mononuclear derived peptides induce IL-10 production in human monocytes.
cells. Life Sci 53: 1735–1742, 1993. J Immunol 156: 2517–2521, 1996.
10. AOKI Y, IWASAKI Y, KATAHIRA M, YOISO, AND SAITO H. Regula- 30. BIGLIARDI PL, BIGLIARDI-QI M, BUECHNER S, AND RUFLI T.
tion of the rat proopiomelanocortin gene expression in AtT-20 Expression of ␮-opiate receptor in human epidermis and keratin-
cells. I. Effects of common segregatogues. Endocrinology 138: ocytes. J Invest Dermatol 111: 297–301, 1998.
1923–1929, 1997. 31. BIRCHALL N, ORLOW SJ, KUPPER T, AND PAWELEK J. Interac-
11. AOKI Y, IWASAKI Y, KATAHIRA M, OISO Y, AND SAITO H. Regu- tions between ultraviolet light and interleukin-1 on MSH binding in
lation of the rat proopiomelanocortin gene expression in AtT-20 both mouse melanoma and human squamous carcinoma cells.
cells. II. Effects of the pituitary adenylate cyclase-activating Biochem Biophys Res Commun 175: 839 – 845, 1991.
polypeptide and vasoactive intestinal polypeptide. Endocrinology 32. BLALOCK JE. Molecular basis for bidirectional communication
138: 1930 –1940, 1997. between the immune and the neuroendocrine systems. Physiol Rev
12. ARENDS RJ, VERMEER H, MARTENS GJ, LEUNISSEN JA, WEN- 69: 1–32, 1989.
DELAAR BONGA SE, AND FLIK G. Cloning and expression of two 33. BLALOCK JE. The syntax of immune-neuroendocrine communica-
proopiomelanocortin mRNAs in the common carp (Cyprinus car- tion. Immunol Today 15: 504 –511, 1994.
pio L.). Mol Cell Endocrinol 143: 23–31, 1998. 34. BOHM M AND LUGER TA. The pilosebaceous unit is part of the skin
13. ASAKURA H, ZWAIN IH, AND YEN SS. Expression of genes encod- immune system. Dermatology 196: 75–79, 1998.
ing corticotropin-releasing factor (CRF), type 1 CRF receptor, and 35. BOHM M, METZE D, SCHULTE U, BECHER E, LUGER TA, AND
CRF-binding protein and localization of the gene products in the BRZOSKA T. Detection of melanocortin-1 receptor antigenicity on
human ovary. J Clin Endocrinol Metab 82: 2720 –2725, 1997. human skin cells in culture and in situ. Exp Dermatol 8: 453– 461,
14. ATWELL WJ. On the nature of the pigmentation changes following 1999.
hypophysectomy in the frog larva. Science 49: 48 –50, 1916. 36. BOHM M, SCHULTE U, KALDEN H, AND LUGER TA. Alpha-mela-
nocyte-stimulating hormone modulates activation of NF-␬B and
15. AUERNHAMMER CJ, CHESNOKOVA V, AND MELMED S. Leuke-
AP-1 and secretion of interleukin-8 in human dermal fibroblasts.
mia inhibitory factor modulates interleukin-1␤-induced activation
Ann NY Acad Sci 885: 277–286, 1999.
of the hypothalamo-pituitary-adrenal axis. Endocrinology 139:
37. BOITANI C, MATHER JP, AND BARDIN CW. Stimulation of adeno-
2201–2208, 1998.
sine 3⬘,5⬘-monophosphate production in rat Sertoli cells by ␣-mela-
16. AUERNHAMMER CJ, CHESNOKOVA V, BOUSQUET C, AND
notropin-stimulating hormone (␣MSH) and des-acetyl ␣MSH. En-
MELMED S. Pituitary corticotroph SOCS-3: novel intracellular reg-
docrinology 118: 1513–1518, 1986.
ulation of leukemia-inhibitory factor-mediated proopiomelanocor-
38. BOLOGNIA JL, MURRAY M, AND PAWELEK J. Evidence that UVB-
tin gene expression and adrenocorticotropin secretion. Mol Endo-
induced melanogenesis is mediated through the MSH receptor
crinol 12: 954 –961, 1998.
system. J Invest Dermatol 92: 651– 660, 1989.
17. AUTELITANO DJ, LUNDBLAD JR, BLUM M, AND ROBERTS JL. 39. BOLOGNIA JL, MURRAY-GONZALES M, AND PAWELEK J. Hairless
Hormonal regulation of POMC gene expression. Annu Rev Physiol pigmented guinea pigs: a new animal model for studies of the
51: 715–726, 1989. pigmentary system. Pigment Cell Res 3: 150 –156, 1990.
18. BAMBERGER CM, WALD M, BAMBERGER AM, ERGÜN S, BEIL 40. BOLOGNIA JL, SODI SA, CHAKRABORTY AK, FARGNOLI MC, AND
FU, AND SCHULTE HM. Human lymphocytes produce urocortin, PAWELEK JM. Effects of ultraviolet irradiation on the cell cycle.
but not corticotropin-releasing hormone. J Clin Endocrinol Metab Pigment Cell Res 7: 320 –325, 1994.
83: 708 –711, 1998. 41. BOMIRSKI A, SLOMINSKI A, AND BIGDA J. The natural history of
19. BARKER DM AND CORDER R. Studies of the role of endothelium- a family of transplantable melanomas in hamsters. Cancer Met Rev
dependent nitric oxide release in the sustained vasodilator effects 7: 95–119, 1988.
of corticotrophin releasing factor and sauvagine. Br J Pharmacol 42. BORRELLI E. A chilled-out knockout. Nature Genet 19: 108 –109,
126: 317–325, 1999. 1998.
20. BATEMAN A, SOLOMON S, AND BENNETT HP. Post-translational 43. BOS JD. Skin Immune System (SIS). Boca Raton FL: CRC, 1997
modification of bovine pro-opiomelanocortin. J Biol Chem 265: 44. BOSTON BA, BLAYDON KM, VARNERIN J, AND CONE RD. Inde-
22130 –22136, 1990. pendent and additive effects of central POMC and leptin pathways
21. BECHER E, MAHNKE K, BHAROWAJ R, SCHWARZ T, AND LUGER on murine obesity. Science 278: 1641–1644, 1997.
TA. Human dendritic cells express a receptor which is specific for 45. BOSTON BA AND CONE RD. Characterization of melanocortin
the immunosuppressing neuropeptide ␣-MSH. J Invest Dermatol receptor subtype expression in murine adipose tissues and in the
109: 254, 1997. 3T3–L1 cell line. Endocrinology 137: 2043–2050, 1996.
22. BECHER E, MAHNKE K, BRZOSKA T, KALDEN DH, GRABBE S, 46. BROUXHON SM, PRASAD AV, JOSEPH SA, FELTEN DL, AND
AND LUGER TA. Human peripheral blood derived dendritic cells BELLINGER DL. Localization of corticotropin-releasing factor in
express functional melanocortin receptor MC-1. Ann NY Acad Sci primary and secondary lymphoid organs of the rat. Brain Behav
885: 188 –195, 1999. Immun 12: 107–122, 1998.
23. BELON PE. UVA exposure and pituitary secretion: variations of 47. BRZOSKA T, KALDEN DH, SCHOLZEN T, AND LUGER TA. Molec-
human lipotropin concentrations (␤LPH) after UVA exposure. Pho- ular basis of the ␣-MSH/IL-1 antagonism. Ann NY Acad Sci 885:
tochem Photobiol 42: 327–329, 1985. 230 –238, 1999.
24. BERNSTEIN JE AND SWIFT R. Relief of intractable pruritus with 48. BRZOSKA T, SCHOLZEN T, BECKER E, HARTMEYER M, BLETZ
naloxone. Arch Dermatol 115: 1366 –1367, 1979. T, SCHWARZ T, AND LUGER TA. UVB irradiation regulates the
25. BERTAGNA X. Proopiomelanocortin-derived peptides. Endocrinol expression of proopiomelanocortin, prohormone convertase 1 and
Metab Clin North Am 23: 467– 485, 1994. melanocortin receptor 1 by human keratinocytes. J Invest Derma-
26. BERTAGNA X, LENNE F, COMAR D, MASSIAS JF, WAJCMAN H, tol 108: 622, 1997.
BAUDIN V, LUTON JP, AND GIRARAD F. Human ␣-melanocyte- 49. BRAR B, SANDERSON T, WANG N, AND LOWRY JP. Post-transla-

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


1010 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

tional processing of human procorticotrophin-releasing factor in man melanocytes and keratinocytes in culture: regulation by UVB.
transfected mouse neuroblastoma and Chinese hamster ovary cell Biochim Biophys Acta 1313: 130 –138, 1996.
lines. J Endocrinol 154: 431– 440, 1997. 68. CHAKRABORTY AK, ORLOW SJ, BOLOGNIA JL, AND PAWELEK J.
50. BRUNETTI L, PREZIOSI P, RAGAZZONI E, AND VACCA M. Involve- Structural/functional relationships between internal and external
ment of nitric oxide in basal and interleukin-1␤-induced CRH and MSH receptors: modulation of expression in Cloudman melanoma
ACTH release in vitro. Life Sci 53: 219 –222, 1993. cells by UVB radiation. J Cell Physiol 147: 1– 6, 1991.
51. BUFFEY J, THODY AJ, BLEEHEN SS, AND S MACNEIL. ␣-Melano- 69. CHAKRABORTY AK, ORLOW SJ, AND PAWELEK JM. Stimulation
cyte-stimulating hormone stimulates protein kinase C activity in of melanocyte-stimulating hormone receptors by retinoic acid.
murine B16 melanoma. J Endocrinol 133: 333–340, 1992. FEBS Lett 276: 205–208, 1990.
52. BUGGY JJ. Binding of ␣-melanocyte-stimulating hormone to its 70. CHAKRABORTY AK AND PAWELEK JM. Upregulation of MSH re-
G-protein-coupled receptor on B-lymphocytes activates the Jak/ ceptor by MSH in Cloudman melanoma cells. Biochem Biophys Res
STAT pathway. Biochem J 331: 211–216, 1998. Commun 188: 1325–1331, 1992.
53. BURCHILL SA AND THODY AJ. Melanocyte-stimulating hormone 71. CHAKRABORTY AK, SLOMINSKI A, ERMAK G, HWANG J, AND
and the regulation of tyrosinase activity in hair follicular melano- PAWELEK JM. Ultraviolet B and melanocyte-stimulating hormone
cytes of the mouse. J Endocrinol 111: 225–232, 1986. (MSH) stimulate mRNA production for ␣MSH receptors and pro-
54. BURCHILL SA, VIRDEN R, FULLER BB, AND THODY AJ. Regula- opiomelanocortin-derived peptides in mouse melanoma cells and
tion of tyrosinase synthesis by ␣-melanocyte-stimulating hormone transformed keratinocytes. J Invest Dermatol 105: 655– 659, 1995.
in hair follicular melanocytes of the mouse. J Endocrinol 116: 72. CHALLIS JRG. CRH, a placental clock and preterm labour. Nature
17–23, 1988. Med 1: 416, 1995.
55. BUZZETTIR R, MCLOUGHLIN L, LAVENDER PM, CLARK AJ, AND 73. CHALMERS DT, LOVENBERG TW, AND DE SOUZA EB. Localiza-
REES LH. Expression of pro-opiomelanocortin gene and quantifi- tion of novel corticotropin-releasing factor receptor (CRF2) mRNA
cation of adrenocorticotropic hormone-like immunoreactivity in expression to specific subcortical nuclei in rat brain: comparison
human normal peripheral mononuclear cells and lymphoid and with CRF1 receptor mRNA expression. J Neurosci 15: 6340 – 6350,
myeloid malignancies. J Clin Invest 83: 733–737, 1989. 1995.
56. CALOGERO AE, BURRELLO N, NEGRI-CESI P, PAPALE L, PU- 74. CHALMERS DT, LOVENBERG TW, GRIGORIADIS DE, BEHAN
LUMBO MA, CIANCI A, SANFILIPPO S, AND D’AGATA R. Effects of DP, AND DE SOUZA EB. Corticotropin-releasing factor receptors:
corticotropin-releasing hormone on ovarian estrogen production in from molecular biology to drug design. Trends Pharmacol Sci 17:
vitro. Endocrinology 137: 4161– 4166, 1996. 166 –172, 1996.
57. CAN G, ABDEL-MALEK Z, PORTER-GILL PA, GILL P, BOYCE S, 75. CHANG AC, CHOCHET M, AND COHEN SN. Structural organization
GRABOWSKI GA, NORDLUND J, AND FAROOQUI J. Identification of human genomic DNA encoding pro-opiomelanocortin peptide.
and sequencing of a putative variant of proopiomelanocortin in Proc Natl Acad Sci USA 77: 4890 – 4894, 1980.
human epidermis and epidermal cells in culture. J Invest Dermatol 76. CHANG AC, ISRAEL A, GAZDAR A, AND COHEN SN. Initiation of
111: 485– 491, 1998. proopiomelanocortin mRNA from a normally quiescent promoter
58. CANNON JG, TATRO JB, REICHLIN S, AND DINARELLO CA. Alpha in a human small cell lung cancer cell line. Gene 84: 115–126, 1989.
melanocyte stimulating hormone inhibits immunostimulatory and 77. CHANG CP, PEARSE RV, O’CONNELL S, AND ROSENFELD MG.
inflammatory actions of interleukin 1. J Immunol 137: 2232–2236, Identification of a seven transmembrane helix receptor for corti-
1986. cotropin-releasing factor and sauvagine in mammalian brain. Neu-
59. CASTRO MG AND MORRISON E. Post-translational processing of ron 11: 1187–1195, 1993.
proopiomelanocortin in the pituitary and in the brain. Crit Rev 78. CHEN R, LEWIS KA, PERRIN MH, AND VALE WW. Expression
Neurobiol 11: 35–57, 1997. cloning of a human corticotropin-releasing-factor receptor. Proc
60. CATANIA A, GERLONI V, PROCACCIA S, AIRAGHI L, MANFREDI Natl Acad Sci USA 90: 8967– 8971, 1993.
MG, LOMATER C, GROSSI L, AND LIPTON JM. The anticytokine 79. CHEN W, KELLY MA, OPITZ-ARAYA X, THOMAS RE, LOW MJ, AND
neuropeptide alpha-melanocyte-stimulating hormone in synovial CONE RD. Exocrine gland dysfunction in MC5-R-deficient mice:
fluid of patients with rheumatic diseases: comparisons with other evidence for coordinated regulation of exocrine gland function by
anticytokine molecules. Neuroimmunomodulation 1: 321–328, melanocortin peptides. Cell 91: 789 –798, 1997.
1994. 80. CHESNOKOVA V, AUERNHAMMER CJ, AND MELMED S. Murine
61. CATANIA A AND LIPTON JM. Alpha-melanocyte stimulating hor- leukemia inhibitory factor gene disruption attenuates the hypo-
mone in the modulation of host reactions. Endocr Rev 14: 564 –576, thalamo-pituitary-adrenal axis stress response. Endocrinology 139:
1993. 2209 –2216, 1998.
62. CATANIA A, MANFREDI MG, AIRAGHI L, VIVIRITO MC, CAPETTI 81. CHEUNG CC, CLIFTON DK, AND STEINER RA. Proopiomelanocor-
A, MILAZZO F, LIPTON JM, AND ZANUSSI C. Plasma concentration tin neurons are direct targets for leptin in the hypothalamus. En-
of cytokine antagonists in patients with HIV infection. Neuroim- docrinology 138: 4489 – 4492, 1997.
munomodulation 1: 42– 49, 1994. 82. CHHAJLANI V AND WIKBERG JE. Molecular cloning and expres-
63. CATANIA A, RAJORA N, CAPSONI F, MINONZIO F, STAR RA, AND sion of the human melanocyte stimulating hormone receptor
LIPTON JM. The neuropeptide alpha-MSH has specific receptors on cDNA. FEBS Lett 309: 417– 420, 1992.
neutrophils and reduces chemotaxis in vitro. Peptides 17: 675– 679, 83. CHIAO H, FOSTER S, THOMAS R, LIPTON J, AND STAR RA.
1996. Alpha-melanocyte-stimulating hormone reduces endotoxin-in-
64. CHAKRABORTY A, BOLOGNIA J, FUNASAKA Y, SLOMINSKI A, duced liver inflammation. J Clin Invest 97: 2038 –2044, 1996.
BOLOGNIA J, SODI S, ICHIHASHI M, AND PAWELEK JM. UV light 84. CLARK AJ, LAVENDER PM, BESSER GM, AND REES LH. Pro-
and MSH receptors. Ann NY Acad Sci 885: 110 –116, 1999. opiomelanocortin mRNA size heterogeneity in ACTH-dependent
65. CHAKRABORTY A AND PAWELEK J. MSH receptors in immortal- Cushing’s syndrome. J Mol Endocrinol 2: 3–9, 1989.
ized human epidermal keratinocytes: a potential mechanism for 85. CLARK AJ, LAVENDER PM, COATES P, JOHNSON MR, AND REES
coordinate regulation of the epidermal-melanin unit. J Cell Physiol LH. In vitro and in vivo analysis of the processing and fate of the
157: 344 –350, 1993. peptide products of the short proopiomelanocortin mRNA. Mol
66. CHAKRABORTY A, FUNASAKA Y, PAWELEK JM, NAGAHAMA M, Endocrinol 4: 1737–1743, 1990.
ITO A, AND ICHIHASHI M. Enhanced expression of melanocortin-1 86. CLIFTON VL, READ MA, LEITCH IM, GILES WB, BOURA AL,
receptor (MC1-r) in normal human keratinocytes during differen- ROBINSON PJ, AND SMITH R. Corticotropin-releasing hormone-
tiation: evidence for increased expression of POMC peptides near induced vasodilatation in the human fetal-placental circulation:
suprabasal layer of epidermis. J Invest Dermatol 112: 853– 860, involvement of the nitric oxide-cyclic guanosine 3⬘,5⬘-monophos-
1999. phate-mediated pathway. J Clin Endocrinol Metab 80: 2888 –2893,
67. CHAKRABORTY A, FUNASAKA Y, SLOMINSKI A, ERMAK G, 1995.
HWANG J, PAWELEK JM, AND ICHIHASHI M. Production and 87. CLIFTON VL, TELFER JF, THOMPSON AJ, CAMERON IT, TEOH
release of proopiomelanocortin (POMC)-derived peptides by hu- TG, LYE SJ, AND CHALLIS JR. Corticotropin-releasing hormone and

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 1011

proopiomelanocortin-derived peptides are present in human myo- late adenosine 3⬘,5⬘-monophosphate production in M2R mouse
metrium. J Clin Endocrinol Metab 83: 3716 –3721, 1998. melanoma cells. Endocrinology 134: 177–185, 1994.
88. CONE RD, LU D, KOPPULA S, VAGE DI, KLUNGLAND H, BOSTON 106. EVANS VR, MANNING AB, BERNARD LH, CHRONWALL BM, AND
B, CHEN W, ORTH DN, POUTON C, AND KESTERSON RA. The MILLINGTON WR. ␣-Melanocyte-stimulating hormone and N-acetyl-
melanocortin receptors: agonists, antagonists, and the hormonal ␤-endorphin immunoreactivities are localized in the human pituitary
control of pigmentation. Recent Prog Horm Res 51: 287–318, 1996. but are not restricted to the zona intermedia. Endocrinology 134:
89. DAI R AND STREILEIN JW. Ultraviolet B-exposed and soluble fac- 97–106, 1994.
tor-pre-incubated epidermal Langerhans cells fail to induce contact 107. FAROOQUI JZ, MEDRANO EE, ABDEL-MALEK Z, AND NORD-
hypersensitivity and promote DNP-specific tolerance. J Invest Der- LUND J. The expression of proopiomelanocortin and various
matol 108: 721–726, 1997. POMC-derived peptides in mouse and human skin. Ann NY Acad
90. DAY R, SCHAFER , WATSON SJ, CHRETIEN M, AND SEIDAH Sci 680: 508 –510, 1993.
MKH/NG. Distribution and regulation of the prohormone conver- 108. FAROOQUI JZ, MEDRANO EE, BOISSY RE, TIGELAAR RE, AND
tases PC1 and PC2 in the rat pituitary. Mol Endocrinol 6: 485– 497, NORDLUND JJ. Thy-1⫹ dendritic cells express truncated form of
1992. POMC mRNA. Exp Dermatol 4: 297–301, 1995.
91. DEBOLD CR, MENEFEE JK, NICHOLSON WE, AND ORTH DN. 109. FAZAL N, SLOMINSKI A, CHOUDHRY MA, WEI ET, AND SAYEED
Proopiomelanocortin gene is expressed in many normal human MM. Effect of CRF and related peptides on calcium signalling in
tissues and in tumors not associated with ectopic adrenocortico- human and rodent melanoma cells. FEBS Lett 435: 187–190, 1998.
tropin syndrome. Mol Endocrinol 2: 862– 870, 1988. 110. FELING P, BAXTER JD, AND FROHMAN LA. Endocrinology and
92. DEGANI H, DEJORDY JO, AND SALOMON Y. Stimulation of cAMP Metabolism (3rd ed.). New York: McGraw-Hill, 1995.
and phosphomonoester production by melanotropin in melanoma 111. FISCHER EG. Opioid peptides modulate immune functions. Im-
cells: 31P NMR studies. Proc Natl Acad Sci USA 88: 1506 –1510, munopharmacol Immunotoxicol 10: 265–326, 1988.
1991. 112. FISHER LA. Central actions of corticotropin-releasing factor on
93. DESCHODT-LANCKMAN M, VANNESTE Y, LOUR B, MICHEL A, autonomic nervous activity and cardiovascular functioning. Ciba
LIBERT A, CHANEM G, AND LEJEUNE F. Degradation of ␣-mela- Found Symp 172: 243–253, 1993.
nocyte stimulating hormone (␣-MSH) by CALLA/endopeptidase 113. FITZPATRICK TB, EISEN AZ, WOLFF K, FREEDBERG IM, AND
24.11 expressed by human melanoma cells in culture. Int J Cancer AUSTEN KF. Dermatology in General Medicine. New York:
46: 1124 –1130, 1990. McGraw Hill, 1987.
94. DIBLASIO AM, GIRALDI FP, VIGANO P, PETRAGLIA F, BIGNALI 114. FLEISHER-BERKOVICH S AND DANON A. Effect of corticotropin-
M, AND CAVAGNINI F. Expression of corticotropin-releasing hor- releasing factor on prostaglandin synthesis in endothelial cells and
mone and its R1 receptor in human endometrial stromal cells. fibroblasts. Endocrinology 136: 4068 – 4072, 1995.
J Clin Endocrinol Metab 82: 1594 –1597, 1997. 115. FLEISHER-BERKOVICH S, RIMON G, AND DANON A. Cortico-
95. DIPASQUALE A, MCGUIRE J, AND VARGA JM. The number of tropin releasing factor modulates interleukin-1-induced prostaglan-
din synthesis in fibroblasts: receptor binding and effects of antag-
receptors for ␣-melanocyte stimulating hormone in Cloudman mel-
onists. Regul Peptides 77: 121–126, 1998.
anoma cells is increased by dibutyryl adenosine 3⬘:5⬘-cyclic mono-
116. FLEISHER-BERKOVICH S, RIMON G, AND DANON A. Modulation
phosphate or cholera toxin. Proc Natl Acad Sci USA 74: 601– 605,
of endothelial prostaglandin synthesis by corticotropin releasing
1977.
factor and antagonists. Eur J Pharmacol 353: 297–302, 1998.
96. DISSANAYAKE NS AND MASON RS. Modulation of skin cell func-
117. FODOR M, SLUITER A, FRANKHUIZZEN-SIEREVOGE A,
tions by transforming growth factor-␤1 and ACTH after ultraviolet
WEIGANT VM, HOOGERHOUT P, WILDT DJ, AND VERSTEEG DH.
irradiation. J Endocrinol 159: 153–163, 1998.
Distribution of Lys-gamma2-melanocyte stimulating hormone (Lys-
97. DONALDSON CJ, SUTTON SW, PERRIN MH, CORRIGAN AZ,
gamma 2-MSH)-like immunoreactivity in neuronal elements in the
LEWIS KA, JERIVIER, VAUGHAN JM, AND VALE WW. Cloning and
brain and peripheral tissues of the rat. Brain Res 731: 182–189,
characterization of human urocortin. Endocrinology 137: 2167– 1996.
2170, 1996. 118. FUKATA J, IMURA H, AND NAKAO K. Cytokines as mediators in the
98. DORIN RI, ZLOCK DW, AND KILPATRICK K. Transcriptional regu- regulation of the hypothalamic-pituitary-adrenocortical function. J
lation of human corticotropin releasing factor gene expression by Endocrinol Invest 17: 141–155, 1994.
cyclic adenosine 3⬘,5⬘-monophosphate: differential effects at prox- 119. FUNASAKA Y, CHAKRABORTY AK, HAYASHI Y, KOMOTO M,
imal and distal promoter elements. Mol Cell Endocrinol 96: 99 –111, OHASHI A, NAGAHAMA M, INOUE Y, PAWELEK J, AND ICHI-
1993. HASHI M. Modulation of melanocyte-stimulating hormone receptor
99. DROUIN J, NEMER M, CHARRON J, GAGNER JP, JEANNOTTE L, expression on normal human melanocytes: evidence for a regula-
SUN YL, THERRIEN M, AND TREMBLAY Y. Tissue specific activity tory role of ultraviolet B, interleukin-1␣, interleukin-1␤, endothe-
of the pro-opiomelanocortin (POMC) gene and repression by glu- lin-1 and tumor necrosis factor-␣. Br J Dermatol 139: 216 –224,
corticoids. Genome 31: 510 –519, 1989. 1998.
100. DUFAU ML, TINAJERO JC, AND FABBRI A. Corticotropin-releasing 120. FUNASAKA Y AND ICHIHASHI M. The effect of ultraviolet B in-
factor: an antireproductive hormone of the testis. FASEB J 7: duced adult T cell leukemia-derived factor/thioredoxin (ADF/TRX)
299 –307, 1993. on survival and growth of human melanocytes. Pigment Cell Res
101. EBERLE AN. The Melanotropins: Chemistry, Physiology and 10: 68 –73, 1997.
Mechanism of Action. New York: Karger, 1988. 121. FUNASAKA Y, SATO H, CHAKRABORTY AK, OHASHI A, CHROU-
102. EHRHART-BORNSTEIN M, HINSON JP, BORNSTEIN SR, AS- SOS GP, AND ICHIHASHI M. Expression of proopiomelanocortin,
CHERBAUM W, AND VINSON GP. Intra-adrenal interactions in the corticotropin-releasing hormone (CRH), and CRH receptor in mel-
regulation of adrenocortical steroidogenesis. Endocr Rev 19: 101– anoma cells, nevus cells, and normal melanocytes. J Invest Der-
143, 1998. matol Symp Proc 4: 105–109, 1999.
103. ERDEN HF, ZWAIN IH, ASAKURA H, AND YEN SS. Corticotropin- 122. FURKERT J, KLUG U, SLOMINSKI A, EICHMULLER S, MEHLIS B,
releasing factor inhibits luteinizing hormone-stimulated P450c17 KERTSCHER U, AND PAUS R. Identification and measurement of
gene expression and androgen production by isolated thecal cells beta-endorphin levels in the skin during induced hair growth in
of human ovarian follicles. J Clin Endocrinol Metab 83: 448 – 452, mice. Biochim Biophys Acta 1336: 315–322, 1997.
1998. 123. GANTZ I, KONDA Y, TASHIRO T, SHIMOTO Y, MIWA H, MUN-
104. ERMAK G AND SLOMINSKI A. Production of POMC, CRH-R1, MC1, ZERT G, WATSON SJ, DEL VALLE J, AND YAMADA T. Molecular
and MC2 receptor mRNA and expression of tyrosinase gene in cloning of a novel melanocortin receptor. J Biol Chem 268: 8246 –
relation to hair cycle and dexamethasone treatment in the C57BL/6 8250, 1993.
mouse skin. J Invest Dermatol 108: 160 –165, 1997. 124. GARCIA-GARCIA L, FUENTES JA, AND MANZANARES J. Differen-
105. ESHEL Y AND SALOMON Y. Calmodulin-binding peptides interfere tial 5-HT-mediated regulation of sites-induced activation of pro-
with melanocyte-stimulating hormone receptor activity and stimu- opiomelanocortin (POMC) gene expression in the anterior and

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


1012 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

intermediate lobe of the pituitary in male rats. Brain Res 772: 145. HARBOUR DV, SMITH EM, AND BLALOCK JE. A novel processing
115–120, 1997. pathway for proopiomelanocortin in lymphocytes: endotoxin in-
125. GEORGALA S, SCHULPIS KH, PAPACONSTANTINOU ED, AND duction of a new prohormone cleaving enzyme. J Neurosci Res 18:
STRATIGOS J. Raised ␤-endorphin serum levels in children with 95–101, 1987.
atopic dermatitis and pruritus. J Dermatol Sci 8: 125–128, 1994. 146. HARBOUR-MCMENAMIN D, SMITH EM, AND BLALOCK JE. Bacte-
126. GESCHWIND II, HUSEBY RA, AND NISHIOKA R. The effect of rial lipopolysaccharide induction of leukocyte-derived cortico-
melanocyte-stimulating hormone on coat color in the mouse. Re- tropin and endorphins. Infect Immun 48: 813– 817, 1985.
cent Prog Horm Res 28: 91–130, 1972. 147. HARTMEYER M, SCHOLZEN T, BECHER E, BHARDWAJ RS,
127. GHANEM G, LIENARD D, HANSON P, LEJEUNE F, AND FRUH- SCHWARZ T, AND LUGER TA. Human dermal microvascular endo-
LING J. Increased serum ␣-melanocyte stimulating hormone (␣- thelial cells express the melanocortin receptor type 1 and produce
MSH) in human malignant melanoma. Eur J Cancer Clin Oncol 22: increased levels of IL-8 upon stimulation with alpha-melanocyte-
535–536, 1986. stimulating hormone. J Immunol 159: 1930 –1937, 1997.
128. GHANEM G, LOIR B, HADLEY M, ABDEL MALEK Z, LIBERT A, 148. HARTWIG AC. Peripheral ␤-endorphin and pain modulation.
DEL MARMOL V, LEJEUNE F, LOZANO J, AND GARCIA-BORRON Anesth Prog 38: 75–78, 1991.
JC. Partial characterization of IR-␣-MSH peptides found in mela- 149. HEDLEY SJ, GAWKRODGER DJ, WEETMAN AP, MORANDINI R,
noma tumors. Peptides 13: 989 –994, 1992. BOEYNAEMS J-M, GHANEM G, AND NEIL SM. ␣-Melanocyte stim-
129. GHANEM G, VERSTEGEN J, DE RIJCKE S, HANSON P, VAN ulating hormone inhibits tumour necrosis factor-␣ stimulated in-
ONDERBERGEN A, LIBERT A, DEL MARMOL V, ARNOULD R, tercellular adhesion molecule-1 expression in normal cutaneous
VERCAMMEN-GRANDJEAN A, AND LEJEUNE F. Studies on fac- human melanocytes and in melanoma cell lines. Br J Dermatol 138:
tors influencing human plasma ␣-MSH. Anticancer Res 9: 1691– 536 –543, 1998.
1696, 1989. 150. HEINRICH N, MEYER MR, FURKERT J, SASSE A, BEYERMANN
130. GHANEM G, COMUNALE G, LIBERT A, VERCAMMEN-GRAND- M, BONIGK W, AND BERGER H. Corticotropin-releasing factor
JEAN A, AND LEJEUNE FJ. Evidence for alpha-melanocyte-stimu- (CRF) agonists stimulate testosterone production in mouse leydig
lating hormone (␣-MSH) receptors on human malignant melanoma cells through CRF receptor-1. Endocrinology 139: 651– 658, 1998.
cells. Int J Cancer 41: 248 –255, 1988. 151. HELDWEIN KA, REDICK DL, RITTENBERG MB, CLAYCOMB WC,
131. GHANEM GE, VERSTEGEN J, LIBERT A, ARNOULD R, AND LE- AND STENZEL-POORE MP. Corticotropin-releasing hormone recep-
JEUNE F. ␣-Melanocyte-stimulating hormone immunoreactivity in tor expression and functional coupling in neonatal cardiac myo-
human melanoma metastases extracts. Pigment Cell Res 2: 519 – cytes and AT-1 cells. Endocrinology 137: 3631–3639, 1996.
523, 1989. 152. HILTZ ME, CATANIA A, AND LIPTON JM. Alpha-MSH peptides
132. GIRALDI FP AND CAVAGNINI F. Corticotropin-releasing hormone inhibit acute inflammation induced in mice by rIL-1 beta, rIL-6,
is produced by rat corticotropes and modulates ACTH secretion in rTNF-alpha and endogenous pyrogen but not that caused by LTB4,
a paracrine/autocrine fashion. J Clin Invest 101: 2478 –2484, 1998. PAF and rIL-8. Cytokine 4: 320 –328, 1992.
133. GIRARD JP AND SPRINGER TA. High endothelial venules (HEVs): 153. HILTZ ME AND LIPTON JM. Alpha-MSH peptides inhibit acute
specialized endothelium for lymphocyte migration. Immunol To- inflammation and contact sensitivity. Peptides 11: 979 –982, 1990.
day 16: 449 – 457, 1995. 154. HINNEY A, BECKER I, HEIBULT O, NOTTEBOM K, SCHMIDT A,
134. GJERDE EA, WOIE KA, WEI ET, AND REED RK. Corticotropin- ZIEGLER A, MAYER H, SIEGFRIED W, BLUM WF, REMSCHMIDT
releasing hormone inhibits lowering of interstitial fluid pressure in H, AND HEBEBRAND J. Systematic mutation screening of the pro-
rat trachea induced by neurogenic inflammation. Eur J Pharmacol opiomelanocortin gene: identification of several genetic variants
352: 99 –102, 1998. including three different insertions, one nonsense and two mis-
135. GLINSKI W, BRODECKA H, GLINSKA-FERENZ M, AND KOWALSKI sense point mutations in probands of different weight extremes.
D. Neuropeptides in psoriasis: possible role of beta-endorphin in J Clin Endocinol Metab 83: 3737–3741, 1998.
the pathomechanism of the disease. Int J Dermatol 33: 356 –361, 155. HIROBE T. Melanocyte stimulating hormone induces the differen-
1994. tiation of mouse epidermal melanocytes in serum-free culture.
136. GLINSKI W, BRODECKA H, GLINSKA-FERENZ M, AND KOWALSKI J Cell Physiol 152: 337–345, 1992.
D. Increased concentration of beta-endorphin in the sera of pa- 156. HOGG N AND BERLIN C. Structure and function of adhesion recep-
tients with severe atopic dermatitis. Acta Derm Venereol 75: 9 –11, tors in leukocyte trafficking. Immunol Today 16: 327–330, 1995.
1995. 157. HOLTZMANN H, ALTMEYER P, AND SCHULTZ-AMLING W. Der
137. GOLDSMITH LA. Physiology, Biochemistry, and Molecular Biol- einfluss ultravioletter strtahlen auf die hypothalamus-hypophysen-
ogy of the Skin. New York: Oxford Univ. Press, 1991. achse des menschen. Acta Dermatol 8: 119 –123, 1982.
138. GRABBE S, BHARDWAJ RS, MAHNKE K, SIMON MM, SCHWARZ 158. HOLTZMANN H, ALTMEYER P, STOHR L, AND CHILF GN. Die
T, AND LUGER TA. Alpha-melanocyte-stimulating hormone induces beemfussung des alpha-MSH durch UVA-bestrahlunger der hautein
hapten-specific tolerance in mice. J Immunol 156: 473– 478, 1996. funktionstest. Hautarzt 34: 294 –297, 1983.
139. GRAMMATOPOULOS D, DAI Y, CHEN J, KARTERIS E, PAPADO- 159. HORSCH D, DAY R, SEIDAH NG, WEIHE E, AND SCHAFER MK.
POULOU N, EASTON AJ, AND HILLHOUSE EW. Human cortico- Immunohistochemical localization of the pro-peptide processing
tropin-releasing hormone receptor: differences in subtype expres- enzymes PC1/PC3 and PC2 in the human anal canal. Peptides 18:
sion between pregnant and nonpregnant myometria. J Clin 755–760, 1997.
Endocrinol Metab 83: 2539 –2544, 1998. 160. HRUBY VJ, WILKES BC, HADLEY ME, AL-OBEIDI F, SAWYER TK,
140. GRANHOLM NH AND VAN AMERONGEN AW. Effects of exoge- STAPLES DJ, DEVAUX AE, DYM O, CASTRUCCI AM, HINTZ MF,
nous MSH on the transformation from phaeo- to eumelanogenesis RIEHM JP, AND RAO KR. ␣-Melanotropin: the minimal active se-
within C57BL/6J-Ay/ a hairbulb melanoctyes. J Invest Dermatol 96: quence in the frog skin bioassay. J Med Chem 30: 2126 –2130, 1987.
78 – 84, 1991. 161. HUNT G. Melanocyte-stimulating hormone: a regulator of human
141. HADLEY M. Endocrinology. Englewood Cliffs, NJ: Prentice Hall, melanocyte physiology. Pathobiology 63: 12–21, 1995.
1996. 162. HYATT P, BELL JB, BHATT K, CHU FW, TAIT JF, TAIT SA, AND
142. HAGAN P, POOLE S, AND BRISTOW AF. Immunosuppressive ac- WHITLEY GS. Effects of ␣-melanocyte-stimulating hormone on the
tivity of corticotrophin-releasing factor. Biochem J 281: 251–254, cyclic AMP and phospholipid metabolism of rat adrenocortical
1992. cells. J Endocrinol 110: 405– 416, 1986.
143. HAKANISHI S, TERANISHI Y, WATANABI Y, NOTAKE M, NODA 163. IKEDA K, TOHO K, SATO S, EBISAWA T, TOKUDOME G, HOSOYA
M, KAKIDANI H, INGAMI HJ, AND NUMA S. Isolation and charac- T, HARADA M, NAKAGAWA O, AND NAKAO K. Urocortin, a newly
terization of the bovine corticotropin/betalipotropin precursor identified corticotropin-releasing factor-related mammalian pep-
gene. Eur J Biochem 115: 429 – 434, 1981. tide, stimulates atrial natriuretic peptide and brain natriuretic pep-
144. HARBOUR DV, GALIN FS, HUGHES TK, SMITH EM, AND BLA- tide secretions from neonatal rat cardiomyocytes. Biochem Bio-
LOCK JE. Role of leukocyte-derived pro-opiomelanocortin pep- phys Res Commun 250: 298 –304, 1998.
tides in endotoxic shock. Circ Shock 35: 181–191, 1991. 164. IM S, MORO O, PENG F, MEDRANO EE, CORNELIUS J, BAB-

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 1013

COCK G, NORDLUND JJ, AND ABDEL-MALEK ZA. Activation of the 184. KAWAHITO Y, SANO H, MUKAI S, ASAI K, KIMURA S,
cyclic AMP pathway by ␣-melanotropin mediates the response of YAMAMURA Y, KATO H, CHROUSOS GP, WILDER RL, AND
human melanocytes to ultraviolet B radiation. Cancer Res 58: 47– KONDO M. Corticotropin releasing hormone in colonic mucosa in
54, 1998. patients with ulcerative colitis. Gut 37: 544 –551, 1995.
165. INUI A. Feeding and body-weight regulation by hypothalamic neu- 185. KEMP CF, WOODS RJ, AND LOWRY PJ. The corticotrophin-releas-
ropeptides: mediation of the actions of leptin. Trends Neurosci 22: ing factor-binding protein: an act of several parts. Peptides 19:
62– 67, 1999. 1119 –1128, 1997.
166. ITOI K, SEASHOLTZ AF, AND WATSON SJ. Cellular and extracel- 186. KIANG JG. Corticotropin-releasing factor increases [Ca2⫹]; via re-
lular regulatory mechanisms of hypothalamic corticotropin-releas- ceptor-mediated Ca2⫹ channels in human epidermoid A-431 cells.
ing hormone neurons. Endocr J 45: 13–33, 1998. Eur J Pharmacol 267: 135–142, 1994.
167. JEANNOTTE L, BURBACH JP, AND DROUIN J. Unusual proopio- 187. KIANG JG. Mystixin-7 and mystixin-11 increase cytosolic free Ca2⫹
melanocortin ribonucleic acids in extra pituitary tissues: intronless and inositol trisphosphates in human A-431 cells. Eur J Pharmacol
transcripts in testes and long poly A tails in hypothalamus. Mol 291: 107–113, 1995.
Endocrinol 1: 749 –757, 1987. 188. KIANG JG. Corticotropin-releasing factor-like peptides increase
168. JOHANSSON O AND LIU PY. ␣-Melanocyte stimulating hormone cytosolic [Ca2⫹] in human epidermoid A-431 cells. Eur J Pharma-
(␣-MSH)-like immunoreactivity is present in certain normal human col 329: 237–244, 1997.
keratinocytes. Exp Dermatol 2: 204 –208, 1993. 189. KIPPENBERGER S, BERNDS A, LOITSCH AS, RAMIREZ-BOSCA
169. JOHANSSON O, LJUNGBERG A, HAN SW, AND VAALASTI A. Evi- A, BEREITER-HAHN J, AND HOLZMANN H. ␣-MSH is expressed in
dence for ␣-melanocyte stimulating hormone containing nerves cultured human melanocytes and keratinocytes. Eur J Dermatol 5:
and neutrophilic granulocytes in the human skin by indirect immu- 395–397, 1995.
nofluorescence. J Invest Dermatol 96: 852– 856, 1991. 190. KISHIMOTO T, PEARSE RV II, LIN CR, AND ROSENFELD MG. A
170. JONES MT AND GILLHAM B. Factors involved in the regulation of sauvagine/corticotropin-releasing factor receptor expressed in
adrenocorticotropic hormone/␤-lipotropic hormone. Physiol Rev heart and skeletal muscle. Proc Natl Acad Sci USA 92: 1108 –1112,
68: 743– 818, 1988. 1995.
171. JORDAN SA AND JACKSON IJ. Melanocortin receptors and antag- 191. KISS M, WLASCHEK M, BRENNEISEN P, MICHEL G, HOMMEL C,
onists regulate pigmentation and body weight. Bioessays 20: 603– LANGE T, PEUS D, KEMENY L, DOBOZY A, SCHARFFETTER-
606, 1998. KOCHANEK K, AND RUZICKA T. ␣-Melanocyte stimulating hor-
172. KALDEN DH, FASTRICH M, BRZOSKA T, SCHOLZEN T, HART- mone induces collagenase/matrix metalloproteinase-1 in human
MEYER M, SCHWARZ T, AND LUBER TA. Alpha-melanocyte-stim- dermal fibroblasts. J Biol Chem 376: 425– 430, 1995.
ulating hormone reduces endotoxin-induced activation of nuclear 192. KJAER A, LARSEN PJ, KNIGGE U, JORGENSEN H, AND WAR-
factor-␬B in endothelial cells. J Invest Dermatol 110: 495, 1998. BERG J. Neuronal histamine and expression of corticotropin-re-
173. KALDEN DH, FASTRICH M, SCHOLZEN T, BRZOSKA T, AND LU- leasing hormone, vasopressin and oxytocin in the hypothalamus:
BER TA. ␣-MSH modulates the expression of LPS-induced adhe- relative importance of H1 and H2 receptors. Eur J Endocrinol 139:
sion molecules. Exp Dermatol 7: 225, 1999. 238 –243, 1998.
174. KALDEN DH, SCHOLZEN T, BRZOSKA T, AND LUGER TA. Mech- 193. KORNER A AND PAWELEK J. Activation of melanoma tyrosinase
anisms of the anti-inflammatory effects of ␣-MSH: role of transcrip- by a cyclic AMP-dependent protein kinase in a cell-free system.
tion factor NF-␬B and adhesion molecule expression. Ann NY Acad Nature 267: 444 – 447, 1977.
Sci 885: 254 –261, 1999. 194. KOSTICH WA, CHEN A, SPERLE K, AND LARGENT BL. Molecular
175. KAMEYAMA K, MONTAGUE PM, AND HEARING VJ. Expression of identification and analysis of a novel human corticotropin-releasing
melanocyte stimulating hormone receptors correlates with mam- factor (CRF) receptor: the CRF2gamma receptor. Mol Endocrinol
malian pigmentation, and can be modulated by interferons. J Cell 12: 1077–1085, 1998.
Physiol 137: 35– 44, 1988. 195. KRAVCHENCO IV AND FURALEV VA. Secretion of immunoreactive
176. KAMEYAMA K, TANAKA S, ISHIDA Y, AND HEARING VJ. Interfer- corticotropin releasing factor and adrenocorticotropic hormone by
ons modulate the expression of hormone receptors on the surface T- and B-lymphocytes in response to cellular stress factors. Bio-
of murine melanoma cells. J Clin Invest 83: 213–221, 1989. chem Biophy Res Commun 204: 828 – 834, 1994.
177. KAPAS S, ORFORD CD, BARKER S, VINSON GP, AND HINSON JP. 196. KRUDE H, BIEBERMANN H, LUCK W, HORN R, BRABANT G, AND
Studies on the intracellular mechanism of action of ␣-melanocyte- GRUTERS A. Severe early-onset obesity, adrenal insufficiency and
stimulating hormone on rat adrenal zona glomerulosa. J Mol En- red hair pigmentation caused by POMC mutations in humans.
docrinol 9: 47–54, 1992. Nature Genet 19: 155–157, 1998.
178. KAPAS S, PURBRICK A, BARKER S, VINSON GP, AND HINSON JP. 197. KUCHLER K AND THORNER J. Secretion of peptides lacking hy-
␣-Melanocyte-stimulating hormone-induced inhibition of angioten- drophobic signal sequence: the role of adenosine triphosphate-
sin II receptor-mediated events in the rat adrenal zona glomerulosa. driven membranes translocators. Endocr Rev 13: 499 –514, 1992.
J Mol Endocrinol 13: 95–104, 1994. 198. KURYSHEV YA, CHILDS GV, AND RITCHIE AK. Corticotropin-re-
179. KAPAS S, PURBRICK A, AND HINSON JP. Role of tyrosine kinase leasing hormone stimulates Ca2⫹ entry though L- and P-type chan-
and protein kinase C in the steroidogenic actions of angiotensin II, nels in rat corticotropes. Endocrinology 137: 2269 –2277, 1996.
␣-melanocyte-stimulating hormone and corticotropin in the rat 199. LAFLAMME N, FEUVRIER E, RICHARD D, AND RIVEST S. Involve-
adrenal cortex. Biochem J 305: 433– 438, 1995. ment of serotonergic pathways in mediating the neuronal activity
180. KARALIS K, SANO H, REDWINE J, LISTWAK S, WILDER RL, AND and genetic transcription of neuroendocrine corticotropin-releas-
CHROUSOS GP. Autocrine or paracrine inflammatory actions of ing factor in the brain of systemically endotoxin-challenged rats.
corticotropin-releasing hormone in vivo. Science 254: 421– 423, Neuroscience 88: 223–240, 1999.
1991. 200. LEITCH IM, BOURA ALA, BOTTI C, READ MA, WALTERS WA, AND
181. KARTERIS E, GRAMMATOPOULOS D, DAI Y, OLAH KB, GHO- SMITH R. Vasodilator actions of urocortin and related peptides in
BARA TB, EASTON A, AND HILLHOUSE EW. The human placenta the human perfused placenta in vitro. J Clin Endocrinol Metab 83:
and fetal membranes express the corticotropin-releasing hormone 4510 – 4513, 1998.
receptor 1␣ (CRH-1␣) and the CRH-C variant receptor. J Clin 201. LERNER AB. The discovery of the melanotropins: a history of
Endocrinol Metab 83: 1376 –1379, 1998. pituitary endocrinology. Ann NY Acad Sci 680: 1–12, 1993.
182. KATAHIRA M, IWASAKI Y, AOKI Y, OISO Y, AND SAITO H. Cyto- 202. LERNER AB AND MCGUIRE J. Effect of alpha- and beta-melanocyte
kine regulation of the rat proopio-melanocortin gene expression in stimulating hormones on the skin color of man. Nature 189: 176 –
AtT-20 cells. Endocrinology 139: 2414 –2422, 1998. 179, 1961.
183. KAVELAARS A, BERKENBOSCH F, CROISET G, REBALLIEUX 203. LEVINE N, LEMUS-WILSON A, WOOD SH, ABDEL-MALEK ZA,
RE, AND HEIJNEN CJ. Induction of ␤-endorphin secretion by lym- AL-OBEIDI F, HRUBY VJ, AND HADLEY ME. Stimulation of follic-
phocytes after subcutaneous administration of corticotropin-re- ular melanogenesis in the mouse by topical and injected mela-
leasing factor. Endocrinology 126: 759 –764, 1990. notropins. J Invest Dermatol 89: 269 –273, 1987.

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


1014 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

204. LEVINE N, SHEFTEL SN, EYTAN T, DORR RT, HADLEY ME, 224. LYONS PD AND BLALOCK JE. The kinetics of ACTH expression in
WEINRACH JC, ERTL GA, TOTH K, MCGEE DL, AND HRUBY VJ. rat leukocyte subpopulations. J Neuroimmunol 63: 103–112, 1995.
Induction of skin tanning by subcutaneous administration of a 225. LYONS PD AND BLALOCK JE. Proopiomelanocortin gene expres-
potent synthetic melanotropin. JAMA 266: 2730 –2736, 1991. sion and protein processing in rat mononuclear leukocytes. J Neu-
205. LEVINS PC, CARR DB, FISHER JE, MOMTAZ K, AND PARRISH JA. roimmunol 78: 47–56, 1997.
␤-Endorphin and ␤-lipotropin response to ultraviolet radiation. 226. MAKRIGIANNAKIS A, MARGIORIS A, MARKOGIANNAKIS E,
Lancet 2: 166, 1983. STOURNARAS C, AND GRAVANIS A. Steroid hormones regulate the
206. LI HE, HEDGER MP, CLEMENTS JA, AND RISBRIDGER GP. Local- release of immunoreactive ␤-endorphin from the Ishikawa human
ization of immunoreactive beta-endorphin and adrenocorticotropic endometrial cell line. J Clin Endocrinol Met 75: 584 –589, 1992.
hormone and pro-opiomelanocortin mRNA to rat testicular inter- 227. MANNA SK AND AGGARHWAL BB. Alpha-melanocyte-stimulating
stitial tissue macrophages. Biol Reprod 45: 282–289, 1991. hormone inhibits the nuclear transcription factor NF-Kappa-B ac-
207. LIAW CW, LOVENBERG TW, BARRY G, OLTERSDORF T, DE- tivation induced by various inflammatory agents. J Immunol 161:
GRIGORIADIS, AND DE SOUZA EB. Cloning and characterization of 2873–2880, 1998.
the human corticotropin-releasing factor-2 receptor complemen- 228. MAURER M, FISCHER E, HANDJISKI B, VON STEBUT E, ALGER-
tary deoxyribonucleic acid. Endocrinology 137: 72–77, 1996. MISSEN B, BAVANDI A, AND PAUS R. Activated skin mast cells
208. LICINIO J, GOLD PW, AND WONG ML. A molecular mechanism for are involved in murine hair follicle regression (catagen). Lab Invest
stress-induced alterations in susceptibility to disease. Lancet 346: 77: 319 –332, 1997.
104 –106, 1995. 229. MAY MJ AND GHOSH S. Signal transduction through NF-kappa B.
209. LIPTON JM AND CATANIA A. Mechanisms of anti-inflammatory Immunol Today 19: 80 – 88, 1998.
action of the neuroimmunomodulatory peptide alpha-MSH. Ann 230. MAZURKIEWICZ JE, CORLISS D, AND SLOMINSKI AT. Differential
NY Acad Sci 840: 373–380, 1998. temporal and spatial expression of POMC mRNA and of the pro-
210. LIPTON JM, CERIANI G, MACALUSO A, MCCOY D, CARNES K, duction of POMC peptides during the murine hair cycle. Ann NY
BILTZ J, AND CATANIA A. Anti-inflammatory effects of the neu- Acad Sci 885: 427– 429, 1999.
ropeptide alpha-MSH in acute, chronic, and systemic inflammation. 231. MAZURKIEWICZ JE, ERMAK G, HWANG J, CHHIAP V, CORLISS
Ann NY Acad Sci 741: 137–148, 1994. D, AND SLOMINSKI A. Towards identification of cells expressing
211. LIU P-Y AND JOHANSSON O. Immunohistochemical evidence of ␣-, POMC in the skin. J Invest Dermatol 104: 636, 1995.
␤-, and ␥3-melanocyte stimulating hormone expression in cutane- 232. MCCUBBIN JA, KAUFFMANN PG, AND NEMEROFF CB. Stress,
ous malignant melanoma of nodular type. J Dermatol Sci 10: Neuropeptides, and Systemic Disease. San Diego, CA: Academic,
203–212, 1998. 1991.
212. LIU P-Y, LONTZ W, BONDESSON L, AND JOHANSSON O. The 233. MCLANE J, OSBER M, AND PAWELEK JM. Phosphorylated isomers
possible role of ␣, ␤ and ␥3 melanocyte stimulating hormone- of L-dopa stimulate MSH binding capacity and responsiveness to
containing keratinocytes in the initiation of vitiligo vulgaris. Eur J MSH in cultured melanoma cells. Biochem Biophys Res Commun
Dermatol 5: 625– 630, 1995.
145: 719 –725, 1987.
213. LOGAN A AND WEATHERHEAD B. Effects of ␣-melanocyte-stimu-
234. MCLANE JA AND PAWELEK JM. Receptors for ␣-MSH in synchro-
lating hormone and [8-arginine] vasotocin upon melanogenesis in
nized Cloudman melanoma cells exhibit positive cooperative in the
vitro. J Endocrinol 91: 501–507, 1981.
late S and G2 phase of the cell cycle. Biochemistry 27: 3743–3747,
214. LOIR B, BOUCHARD B, MORANDINI R, DEL MARMOL V, DERAE-
1988.
MAECKER R, GARCIA-BORRON JC, AND GHANEM G. Immunore-
235. MCLEAN M, BISITS A, DAVIES J, WOODS R, LOWRY P, AND SMITH
active ␣-melanotropin as an autocrine effector in human melanoma
R. A placental clock controlling the length of human pregnancy.
cells. Eur J Biochem 244: 923–930, 1997.
Nature Med 1: 460 – 463, 1995.
215. LOIR B, SALES F, DERAEMAECKER R, MORANDINI R, GARCIA-
236. MCLEOD SD, SMITH C, AND MASON RS. Stimulation of tyrosinase
BORRON JC, AND GHANEM G. ␣-Melanotropin immunoreactivity
in human melanoma exudate is related to necrosis. Eur J Cancer in human melanocytes by pro-opiomelanocortin-derived peptides.
34: 424 – 426, 1998. J Endocrinol 146: 439 – 447, 1995.
216. LOLAIT SJ, CLEMENTS JA, MARKWICK AJ, CHENG C, MCNALLY 237. MCLOON LK AND WIRTSCHAFTER J. Local injections of cortico-
M, SMITH A, AND FUNDER JW. Pro-opiomelanocortin messenger tropin releasing factor reduce doxorubicin-induced acute inflam-
ribonucleic acid and posttranslational processing of beta endor- mation in the eyelid. Invest Ophthalmol Visual Sci 38: 834 – 841,
phin in spleen macrophages. J Clin Invest 77: 1776 –1779, 1986. 1997.
217. LOVENBERG TW, LIAW CW, GRIGORIADIS DE, CLEVENGER W, 238. MECHANICK JI, LEVIN N, ROBERTS JL, AND AUTELITANO DJ.
CHALMERS DT, DE SOUZA EB, AND OLTERSDORF T. Cloning and Proopiomelanocortin gene expression in a distinct population of
characterization of a functionally distinct corticotropin-releasing rat spleen and lung leukocytes. Endocrinology 131: 518 –525, 1992.
factor receptor subtype from rat brain. Proc Natl Acad Sci USA 92: 239. MELZIG MF. Corticotropin releasing factor inhibits proliferation of
836 – 840, 1995. AtT-20 cells. In Vitro Cell Dev Biol 30: 741–743, 1994.
218. LOWRY PJ, WOODS RJ, AND BAIGENT S. Corticotropin releasing 240. MEYER WJ III, SMITH EM, RICHARDS GE, CAVALLO A, MORRILL
factor and its binding protein. Pharmacol Biochem Behav 54: 305– AC, AND BLALOCK JE. In vivo immunoreactive adrenocorticotropin
308, 1996. (ACTH) production by human mononuclear leukocytes from nor-
219. LUGER TA, SCHAUER E, TRAUTINGER F, KRUTMANN J, ANSEL mal and ACTH-deficient individuals. J Clin Endocrinol Metab 64:
J, SCHWARZ A, AND SCHWARZ T. Production of immunosuppress- 98 –105, 1987.
ing melanotropins by human keratinocytes. Ann NY Acad Sci 680: 241. MILLINGTON WR, EVANS VR, BATTIE CN, BAGASRA O, AND
567–570, 1993. FORMAN LJ. Proopiomelanocortin-derived peptides and mRNA
220. LUGER TA, SCHOLZEN T, BRZOSKA T, BECHER E, SLOMINSKI are expressed in rat heart. Ann NY Acad Sci 680: 575–578, 1993.
A, AND PAUS R. Cutaneous immunomodulation and coordination of 242. MIYATA I, SHIOTA C, IKEDA Y, OSHIDA Y, CHAKI S, OKUYAMA
skin stress responses by alpha-melanocyte-stimulating hormone. A, AND INAGAMI T. Cloning and characterization of a short variant
Ann NY Acad Sci 840: 381–394, 1998. of the corticotropin-releasing factor receptor subtype from rat
221. LUGER TA, SCHWARZ T, KALDEN H, SCHOLZEN T, SCHWARZ A, amgdala. Biochem Biophys Res Commun 24: 692– 696, 1999.
AND BRZOSKA T. Role of epidermal cell derived ␣-melanocyte 243. MIZUNO TM, KLEOPOULOS SP, BERGEN TH, ROBERTS JL,
stimulating hormone in ultraviolet light mediated local immuno- PRIEST CA, AND MOBBS CV. Hypothalamic proopio-melanocortin
suppression. Ann NY Acad Sci 885: 209 –216, 1999. mRNA is reduced by fasting and (corrected) in ob/ob and db/db
222. LUNDBLAD JR AND ROBERTS JL. Regulation of propiomelanocor- mice, but is stimulated by leptin. Diabetes 47: 294 –297, 1998.
tin gene expression in pituitary. Endocr Rev 9: 135–158, 1988. 244. MOELLMANN G, SLOMINSKI A, KUKLINSKA E, AND LERNER AB.
223. LUNEC J, PIERON JC, SHERBET GV, AND THODY AJ. Alpha- Regulation of melanogenesis in melanocytes. Pigment Cell Res
melanocortin-stimulating hormone immunoreactivity in melanoma 79 – 87, 1988.
cell. Pathobiology 58: 193–197, 1990. 245. MOODY TW, ZIA F, AND VENUGOPAL R. Corticotropin-releasing

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 1015

factor stimulates cyclic AMP, arachidonic acid release, and growth 70 expression in keratinocytes. J Invest Dermatol 108: 401– 405,
of lung cancer cells. Peptides 15: 281–285, 1994. 1997.
246. MORANDINI R, BOEYNAEMS JM, HEDLEY SJ, MACNEIL S, AND 265. ORLOW SJ, CHAKRABORTY AK, AND PAWELEK JM. Retinoic acid
GHANEM G. Modulation of ICAM-1 expression by alpha-MSH in is a potent inhibitor of inducible pigmentation in mouse and ham-
human melanoma cells and melanocytes. J Cell Physiol 175: 276 – ster melanoma cells. J Invest Dermatol 94: 461– 464, 1990.
282, 1998. 266. ORLOW SJ, HOTCHKISS S, AND PAWELEK JM. Internal binding
247. MORHENN VB. The physiology of scratching: involvement of pro- sites for MSH in wild type and variant Cloudman melanoma cells.
opiomelanocortin gene-coded proteins in Langerhans cells. Prog J Cell Physiol 142: 129 –136, 1990.
Neurol Endocrinol Immunol 4: 265–267, 1991. 267. ORTH DN. Corticotropin-releasing hormone in humans. Endocr
248. MOUNTJOY KG. The human melanocyte stimulating hormone re- Rev 13: 164 –191, 1992.
ceptor has evolved to become “super-sensitive” to melanocortin 268. OWENS A AND NEMEROFF CB. Physiology and pharmacology of
peptides. Mol Cell Endocrinol 102: R7–R11, 1994. corticotropin releasing factor. Pharmacol Rev 43: 425– 473, 1991.
249. MOUNTJOY KG, BIRD IM, RAINEY WE, AND CONE RD. ACTH 268a.PAEZ PEREDA M, SAUER J, PEREZ CASTRO C, FINKIELMAN S,
induces up-regulation of ACTH receptor mRNA in mouse and hu- STALLA GK, HOLSBOER F, AND ARZT E. Corticotropin-releasing
man adrenocortical cell lines. Mol Cell Endocrinol 99: R17–R20, hormone differentially modulates the interleukin-1 system accord-
1994. ing to the level of monocyte activation by endotoxin. Endocrinol-
250. MOUNTJOY KG, ROBBINS LS, MORTRUD MT, AND CONE RD. The ogy 136: 5504 –5510, 1995.
cloning of a family of genes that encode the melanocortin recep- 269. PANERAI AE AND SACERDOTE P. Beta-endorphin in the immune
tors. Science 257: 1248 –1251, 1992. system: a role at last? Immunol Today 18: 317–319, 1997.
251. MOUNTJOY KG AND WONG J. Obesity, diabetes and functions for 270. PAQUET L, ZHOU A, CHANG EY, AND MAINS RE. Peptide biosyn-
proopiomelanocortin-derived peptides. Mol Cell Endocrinol 128: thetic processing: distinguishing prohormone convertases PC1 and
171–177, 1997. PC2. Mol Cell Endocrinol 120: 161–168, 1996.
252. MOZZANICA N, VILLA ML, FOPPA S, VIGNATI G, CATTANEO A, 271. PARK HY, RUSSAKOVSKY V, AO Y, FERNANDEZ E, AND GIL-
DIOTTI R, AND FINZI AF. Plasma alpha-melanocyte-stimulating CHREST BA. ␣-Melanocyte stimulating hormone-induced pigmen-
hormone, beta-endorphin, met-enkephalin, and natural killer cell tation is blocked by depletion of protein kinase C. Exp Cell Res 227:
activity in vitiligo. J Am Acad Dermatol 26: 693–700, 1992. 70 –79, 1996.
253. MUGLIA LJ, JENKINS NA, GILBERT DJ, COPELAND NG, AND 272. PAUS R. Control of the hair cycle and hair diseases as cycling
MAJZOUB JA. Expression of the mouse corticotropin-releasing disorders. Curr Opin Dermatol 3: 248 –258, 1996.
hormone gene in vivo and targeted inactivation in embryonic stem 273. PAUS R, BOTCHKAREV VA, BOTCHKAREVA NV, FECHNER K,
cells. J Clin Invest 93: 2066 –2072, 1994. FURKERT J, ROLOFF B, AND SLOMINSKI A. The skin POMC
254. MUGRIDGE KG, PERRETTI M, GHIARA P, AND PARENTE L. Al- system (SPS): leads and lessons from the hair follicle. Ann NY Acad
pha-melanocyte-stimulating hormone reduces interleukin-1 beta ef- Sci 885: 350 –363, 1999.
274. PAUS R, HANDJISKI B, CZARNETZKI BM, AND EICHMULLER S. A
fects on rat stomach preparations possibly through interference
murine model for inducing and manipulating hair follicle regres-
with a type I receptor. Eur J Pharmacol 197: 151–155, 1991.
sion (catagen): effects of dexamethasone and cyclosporin A. J In-
255. NAGAHAMA M, FUNASAKA Y, FERNANDEZ-FREZ ML, OHASHI
vest Dermatol 103: 143–147, 1994.
A, CHAKRABORTY AK, UEDA M, AND ICHIHASHI M. Immunore-
275. PAUS R, MAURER M, SLOMINSKI A, AND CZARNETZKI BM. Mast
activity of ␣-melanocyte-stimulating hormone, adrenocorticotro-
cell involvement in murine hair growth. Dev Biol 163: 230 –240,
phic hormone and ␤-endorphin in cutaneous malignant melanoma
1994.
and benign melanocytic naevi. Br J Dermatol 138: 981–985, 1998.
276. PAUS R, PETERS EM, EICHMUELLER S, AND BOTCHKAREV VA.
256. NAVILLE D, BARJHOUX L, JAILLARD C, SAEZ JM, DURAND P,
Neural mechanisms of hair growth control. J Invest Dermatol
AND BEGEOT M. Stable expression of normal and mutant human
Symp Proc 2: 61– 68, 1998.
ACTH receptor study of ACTH binding and coupling to adenylate 277. PAWELEK JM. A cyclic AMP requirement for proliferation of
cyclase. Mol Cell Endocrinol 129: 83–90, 1997. Cloudman S91 melanoma cells. Pigment Cell Res 4: 167–176, 1979.
257. NORDLUND JJ, BOISSY RE, HEARING VJ, KING RA, AND POR- 278. PAWELEK JM. Evidence suggesting that a cyclic AMP-dependent
TONNE J. The Pigmentary System: Physiology and Pathophysi- protein kinase is a positive regulator of proliferation in Cloudman
ology. New York: Oxford Univ. Press, 1998. S91 melanoma cells. J Cell Physiol 98: 619 – 625, 1979.
258. OATES EL, ALLAWAY GP, ARMSTRONG GR, BOYAJIAN RA, 279. PAWELEK JM. Studies on the Cloudman melanoma cell line as a
KEHRL JH, AND PRABHAKAR BS. Human lymphocytes produce model for the action of MSH. Yale J Biol Med 58: 571–578, 1985.
pro-opiomelanocortin gene-related transcripts. Effects of lympho- 280. PAWELEK JM. Is human melanogenesis stimulated by cyclic AMP?
tropic viruses. J Biol Chem 263: 10041–10044, 1988. J Invest Dermatol 94: 499 –500, 1990.
259. OATES EL, ALLAWAY GP, AND PRABHAKAR BS. A potential for 281. PAWELEK JM. After dopachrome? Pigment Cell Res 4: 53– 62,
mutual regulation of proopiomelanocortin gene and Epstein-Barr 1991.
virus expression in human lymphocytes. Ann NY Acad Sci 594: 282. PAWELEK JM. Proopiomelanocortin in skin: new possibilities for
60 – 65, 1990. regulation of skin physiology. J Lab Clin Med 122: 627– 628, 1993.
260. ODONOHUE TL AND DORSA DM. The opiomelanotropinergic neu- 283. PAWELEK JM, CHAKRABORTY AL, OSBER MP, ORLOW SJ, MIN
ronal and endocrine system. Peptides 3: 353–395, 1982. KK, ROSENZWEIG KE, AND BOLOGNIA JL. Molecular cascades in
261. OHMORI N, ITOI K, TOZAWA F, SAKAI Y, SAKAI K, HORIBA N, UV-induced melanogenesis: a central role for melanotropins? Pig-
DEMURA H, AND SUDA T. Effect of acetylcholine on corticotrop- ment Cell Res 5: 348 –356, 1992.
ine-releasing factor gene expression in the hypothalamic paraven- 284. PAWELEK JM, FLEISCHMANN R, MCLANE J, GUILLETTE B,
tricular nucleus of conscious rats. Endocrinology 136: 4858 – 4863, EMANUEL J, KORNER A, BERGSTROM A, AND MURRAY M. Stud-
1995. ies on growth and pigmentation of Cloudman S91 melanoma cells.
262. OKI Y, IWABUCHI M, MASUZAWA M, WATANABE F, OZAWA M, Pigment Cell Res 6: 521–533, 1984.
IINO K, TOMINAGA T, AND YOSHIMI T. Distribution and concen- 285. PAWELEK JM, HALABAN R, AND CHRISTIE G. Melanoma cells
tration of urocortin and effect of adrenalectomy on its content in with a cyclic AMP growth requirement. Nature 258: 539 –540, 1975.
rat hypothalamus. Life Sci 62: 807– 812, 1998. 286. PAWELEK JM, KORNER A, BERGSTROM A, AND BOLOGNIA J.
263. OKOSI A, BRAR BK, CHAN M, SOUZA LDE, SMITH E, STEPHA- New regulators of melanin biosynthesis and the autodestruction of
NOU A, LATCHMAN DS, CHOWDREY HS, AND KNIGHT RA. Ex- melanoma cells. Nature 286: 617– 619, 1980.
pression and protective effects of urocortin in cardiac myocytes. 287. PAWELEK JM, SANSONE M, KOCH N, CHRISTIE G, HALABAN R,
Neuropeptides 32: 167–171, 1998. HENDEE J, LERNER AB, AND VARGA JM. Melanoma cells resistant
264. OREL L, SIMON MM, KARLSEDER J, BHARDWAJ R, TRAUT- to inhibition of growth by melanocyte-stimulating hormone. Proc
INGER F, SCHWARZ T, AND LUGER TA. ␣-Melanocyte-stimulating Natl Acad Sci USA 72: 951–956, 1975.
hormone down-regulates differentiation-driven heat shock protein 288. PAWELEK JM, SANSONE M, MOROWITZ J, MOELLMANN G, AND

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


1016 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

GODAWSKA E. Genetic control of melanization: isolation and anal- LEWIS K, SAWCHENKO PE, AND VALE W. Distribution of cortico-
ysis of amelanotic variants from cultured melanotic melanoma tropin-releasing factor receptor mRNA expression in the rat brain
cells. Proc Natl Acad Sci USA 71: 1073–1077, 1974. and pituitary. Proc Natl Acad Sci USA 91: 8777– 8781, 1994.
289. PEARS JS, JUNG RT, BARTLETT W, BROWNING MC, KENICER K, 309. POZZOLI G, BILEZIKJIAN LM, PERRIN MH, BLOUNT AL, AND
AND THODY AJ. A case of skin hyperpigmentation due to ␣-MSH VALE WW. Corticotropin-releasing factor (CRF) and glucocorti-
hypersecretion. Br J Dermatol 126: 286 –289, 1992. coids modulate the expression of type 1 CRF receptor messenger
290. PENG PJ, SAHM UG, DOHERTY RV, KINSMAN RG, MOSS SH, AND ribonucleic acid in rat anterior pituitary cell cultures. Endocrinol-
POUTON CW. Binding and biological activity of C-terminally mod- ogy 137: 65–71, 1996.
ified melanocortin peptides: a comparison between their actions at 310. PRIMUS RJ, YEVICH E, BALTAZAR C, AND GALLAGER DW. Auto-
rodent MC1 and MC3 receptors. Peptides 7: 1001–1008, 1997. radiographic localization of CRF1 and CRF2 binding sites in adult
291. PENHOAT A, JAILLARD C, AND SAEZ JM. Corticotropin positively rat brain. Neuropsychopharmacology 17: 308 –316, 1997.
regulates its own receptors and cAMP response in cultured bovine 311. QUILLAN JM AND SADEE W. Dynorphin peptides: antagonists of
adrenal cells. Proc Natl Acad Sci USA 86: 4978 – 4981, 1989. melanocortin receptors. Pharm Res 14: 713–719, 1997.
293. PERKINS AV, WOLFE CD, EBEN F, SOOTHILL P, AND LINTON EA. 312. RAJORA N, CERIANI G, CATANIA A, STAR RA, MURPHY MT, AND
Corticotropin-releasing hormone-binding protein in human fetal LIPTON JM. alpha-MSH production, receptors, and influence on
plasma. J Endocrinol 146: 395– 401, 1995. neopterin in a human monocyte/macrophage cell line. J Leukoc
294. PERONE MJ, MURRAY CA, BROWN OA, GIBSON S, WHITE A, Biol 59: 248 –253, 1996.
LINTON EA, PERKINS AV, LOWENSTEIN PR, AND CASTRO MG. 313. RAY D AND MELMED S. Pituitary cytokine and growth factor
Procorticotropin-releasing hormone: endoproteolytic processing expression and action. Endocr Rev 18: 206 –228, 1997.
and differential release of its derived peptides within AtT20 cells. 314. REDONDO P, GARCIA-FONCILLAS J, OKROUJNOV I, AND BAN-
Mol Cell Endocrinol 142: 191–202, 1998. DRES E. Alpha-MSH regulates interleukin-10 expression by human
295. PERONE M, WINDEATT S, MORRISON E, SHERING A, TOMASEC keratinocytes. Arch Dermatol Res 290: 425– 428, 1998.
P, LINTON E, LOWENSTEIN P, AND CASTRO M. Intracellular re- 315. RHEINS LA, COTLEUR AL, KLEIER RS, HOPPENJANS WB, SAUN-
tention of the corticotropin-releasing hormone (CRH) precursors DER DN, AND NORDLUND JJ. Alpha-melanocyte stimulating hor-
within COS-7 cells. J Histochem Cytochem 46: 1193–1197, 1998. mone modulates contact hypersensitivity responsiveness in
296. PERRIN M, DONALDSON C, CHEN R, BLOUNT A, BERGGREN T, C57BL6 mice. J Invest Dermatol 93: 511–517, 1989.
BILEZIKJIAN L, SAWCHENKO P, AND VALE W. Identification of a 316. RICHTER RM AND MULVANY MJ. Comparison of hCRF and oCRF
second corticotropin-releasing factor receptor gene and character- effects on cardiovascular responses after central, peripheral, and in
ization of a cDNA expressed in heart. Proc Natl Acad Sci USA 92: vitro application. Peptides 16: 843– 849, 1995.
2969 –2973, 1995. 317. ROBBINS LS, NADEAU JH, JOHNSON KR, KELLY MA, ROSELLI-
297. PETRAGLIA F, FLORIO P, NAPPI C, AND GENAZZANI AR. Peptide REHFUSS L, EBAACK, MOUNTJOY KG, AND CONE RD. Pigmenta-
signaling in human placenta and membranes: autocrine, paracrine, tion phenotypes of variant extension locus alleles result from point
and endocrine mechanisms. Endocr Rev 17: 156 –186, 1996. mutations that alter MSH receptor function. Cell 72: 827– 834, 1993.
298. PETRAGLIA F, FLORIA P, GALLO R, SIMONCINI T, SAVIOZZI M, 318. RODRÍGUEZ-LIÑARES B, LINTON EA, AND PHANEUF S. Expres-
DI BLASIO AM, VAUGHAN J, AND VALE W. Human placenta and sion of corticotrophin-releasing hormone receptor mRNA a protein
fetal membranes express human urocortin mRNA and peptide. in the human myometrium. J Endocrinol 156: 15–21, 1998.
J Clin Endocrinol Metab 81: 3807–3810, 1996. 319. ROE CM, LEITCH IM, BOURA AL, AND SMITH R. Nitric oxide
299. PETRAGLIA F, GARUTI GC, DE RAMUNDO B, ANGIONI S, regulation of corticotropin-releasing hormone release from the
GENAZZANI AR, AND BILEZIKJIAN LM. Mechanism of action of human perfused placenta in vitro. J Clin Endocrinol Metab 81:
interleukin-1␤ in increasing corticotropin-releasing factor and ad- 763–769, 1996.
renocorticotropin hormone release from cultured human placental 320. ROHDE E, FURDERT J, FECHNER K, BEYERMANN M, MULVANY
cells. Am J Obstet Gynecol 163: 1307–1312, 1990. MJ, RICHTER RM, DENEF C, BIENERT M, AND BERGER H. Cor-
300. PETRAGLIA F, POTTER E, CAMERON VA, SUTTON S, BEHAN ticotropin releasing hormone (CRH) receptors in the mesenteric
DP, WOODS RJ, SAWCHENKO PE, LOWRY PJ, AND VALE W. small arteries of rats resemble the (2)-subtype. Biochem Pharma-
Corticotropin-releasing factor-binding protein is produced by hu- col 52: 829 – 833, 1996.
man placenta and intrauterine tissues. J Clin Endocrinol Metab 77: 321. ROLOFF B, FECHNER K, SLOMINSKI A, FURKERT J, BOTCH-
919 –924, 1993. KAREV VA, BULFONE-PAUS S, ZIPPER J, KRAUSE E, AND PAUS
301. PETRAGLIA F, SANTUZ M, FLORIO P, SIMONCINI T, LUISI S, R. Hair cycle-dependent expression of corticotropin releasing fac-
PLAINO L, GENASSANI AR, GENAZZANI AD, AND VOLPE A. Para- tor (CRF) and CRF receptors (CRF-R) in murine skin. FASEB J 12:
crine regulation of human placenta: control of hormonogenesis. J 287–297, 1998.
Reprod Immunol 39: 221–233, 1998. 322. ROSE J. Bilateral adrenalectony induces early onset of summer fur
302. PETRAGLIA F, SAWCHENKO PE, RIVIER J, AND VALE W. Evi- growth in mink (Mustela vison). Comp Biochem Physiol 111: 243–
dence for local stimulation of ACTH secretion by corticotropin- 247, 1995.
releasing factor in human placenta. Nature 328: 717–719, 1987. 323. ROSE J. Adrenocorticotropic hormone (ACTH) but not alpha-mel-
303. PETRAGLIA F, TABANELLI S, GALASSI MC, GARUTI GC, MAN- anocyte stimulating hormone (␣-MSH) as a mediator of adrenalec-
CINI AC, GENAZZANI AR, AND GURPIDE E. Human decidua and in tomy induced hair growth in mink. J Invest Dermatol 110: 456 – 457,
vitro decidualized endometrial stromal cells at term contain immu- 1998.
noreactive corticotropin-releasing factor (CRF) and CRF messen- 324. ROSE J AND STERNER M. The role of the adrenal glands in regu-
ger ribonucleic acid. J Clin Endocrinol Metab 74: 1427–1431, 1992. lating onset of winter fur growth in mink (Mustela vison). J Exp
304. PFROMMER C, KUNZ C, LUGER TA, AND HENZ BM. Proopiomel- Zool 262: 469 – 473, 1992.
anocortin products enhance histamine and LTC4 release from the 325. RUST CC AND MEYER RK. Effect of pituitary autografts on hair
human cell line HMC 1 and human cutaneous mast cells in vitro. color in the short-tailed weasel. Gen Comp Endocrinol 11: 548 –551,
J Invest Dermatol 107: 507, 1996. 1968.
305. PLOTNIKOFF NP, FAITH RE, MURGO AJ, AND GOOD RA. Cyto- 326. RUST CC, SHACKELFORD RM, AND MEYER RK. Hormonal control
kines: Stress and Immunity. Boca Raton, FL: CRC, 1999. of pelage cycles in the mink. J Mammal 46: 549 –565, 1965.
306. POMERANTZ SH AND HUANG LC. Effects of ␤-MSH, cortisol and 327. SAHU A. Evidence suggesting that galanin (GAL), melanin-concen-
ACTH on tyrosinase in the skin of newborn hamsters and mice. trating hormone (MCH), neurotensin (NT), proopiomelanocortin
Endocrinology 87: 302–310, 1970. (POMC) and neuropeptide Y(NPY) are targets of leptin signalling in
307. POTTER E, BEHAN DP, FISCHER WH, LINTON EA, LOWRY PJ, the hypothalamus. Endocr Rev 139: 795–798, 1998.
AND VALE WW. Cloning and characterization of the cDNA for 328. SAHU SN, EDWARDS-PRASAD J, AND PRASAD KN. Alpha tocoph-
human and rat corticotropin releasing factor-binding proteins. Na- erol succinate inhibits melanocyte-stimulating hormone (MSH)-
ture 349: 423– 426, 1991. sensitive adenylate cyclase activity in melanoma cells. J Cell
308. POTTER E, SUTTON S, DONALDSON C, CHEN R, PERRIN M, Physiol 133: 585–589, 1987.

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 1017

329. SAIARDI A AND BORRELLI E. Absence of dopaminergic control on 348. SHIBAHARA S, MORIMOTO Y, FURUTANI Y, NOTAKE M, TAKA-
melanotrophs leads to Cushing’s-like syndrome in mice. Mol En- HASHI H, SHIMIZU S, HORIKAWA S, AND NUMA S. Isolation and
docrinol 12: 1133–1139, 1998. sequence analysis of the human corticotropin-releasing factor pre-
330. SAKAI K, YAMADA M, HORIBA N, WAKUI M, DEMURA H, AND cursor gene. EMBO J 2: 775–779, 1983.
SUDA T. The genomic organization of the human corticotropin- 349. SHIMIZU T AND STREILEIN JW. Influence of alpha-melanocyte
releasing factor type-1 receptor. Gene 219: 125–130, 1998. stimulating hormone on induction of contact hypersensitivity and
331. SAPERSTEIN A, BRAND H, AUDHYA T, NABRISKI D, HUTCHIN- tolerance. J Dermatol Sci 8: 187–193, 1994.
SON B, ROSENZWEIG S, AND HOLLANDER CS. Interleukin 1␤ 350. SHIZUME K, LERNER AB, AND FITZPATRICK TB. In vitro bioassay
mediates stress-induced immunosuppression via corticotropin-re- for the melanocyte stimulating hormone. Endocrinology 54: 553–
leasing factor. Endocrinology 130: 152–158, 1992. 560, 1954.
332. SCATENA C AND ADLER S. Characterization of a human-specific 351. SIEGRIST W AND EBERLE AN. Homologous regulation of the MSH
regulator of placental corticotropin-releasing hormone. Mol Endo- receptor in melanoma cells. J Receptor Res 13: 263–281, 1993.
crinol 12: 1228 –1240, 1998. 352. SIEGRIST W AND EBERLE AN. Melanocortin and their implication
333. SCHAFER M, CARTER L, AND STEIN C. Interleukin 1␤ and corti- in melanoma. Trends Endocrinol Metab 6: 115–120, 1995.
cotropin-releasing factor inhibit pain by releasing opioids from 353. SLOMINSKI A. POMC gene expression in hamster and mouse
immune cells in inflamed tissue. Pharmacology 91: 4219 – 4223, melanoma cells. FEBS Lett 291: 165–168, 1991.
1994. 354. SLOMINSKI A. Identification of ␤-endorphin, ␣-MSH and ACTH
334. SCHAFER M, MOUSA SA, AND STEIN C. Corticotropin-releasing peptides in cultured human melanocytes, melanoma and squamous
factor in antinociception and inflammation. Eur J Pharmacol 323: cell carcinoma cells by RP-HPLC. Exp Dermatol 7: 213–216, 1998.
1–10, 1997. 355. SLOMINSKI A, BAKER J, ERMAK G, CHAKRABORTY A, AND
335. SCHAFER M, MOUSA SA, ZHANG Q, CARTER L, AND STEIN C. PAWELEK J. Ultraviolet B stimulates production of corticotropin
Expression of corticotropin-releasing factor in inflamed tissue is releasing factor (CRF) by human melanocytes. FEBS Lett 399:
required for intrinsic peripheral opioid analgesia. Proc Natl Acad 175–176, 1996.
Sci USA 93: 6096 – 6100, 1996. 356. SLOMINSKI A, BOTCHKAREV NV, BOTCHKAREV VA,
336. SCHALLREUTER K, SLOMINSKI A, PAWELEK JM, JIMBOW K, AND CHAKRABORTY A, LUGER T, UENALAN M, AND PAUS R. Hair
GILCHREST BA. What control melanogenesis? Exp Dermatol 7: cycle dependent production of ACTH in mouse skin. Biochim
143–150, 1998. Biophys Acta 1448: 147–152, 1998.
337. SCHAUER E, TRAUTINGER F, KOCK A, SCHWARZ A, BHARDWAI 357. SLOMINSKI A, BOTCHKAREV V, CHOUDHRY M, FAZAL N, FECH-
R, ANSEL JC, SCHWARZ T, AND LUGER TA. Proopiomelanocortin NER K, FURKERT J, KRAUSE E, ROLOFF B, SAYEED M, WEI E,
derived peptides are synthesized and released by human keratino- ZBYTEK B, ZIPPER J, WORTSMAN J, AND PAUS R. Cutaneous
cytes. J Clin Invest 93: 2258 –2262, 1994. expression of CRH and CRH-R: is there a “skin stress response
338. SCHILLING L, KANZLER C, SCHMIEDEK P, AND EHRENREICH H. system?”. Ann NY Acad Sci 885: 287–311, 1999.
Characterization of the relaxant action of urocortin, a new peptide 358. SLOMINSKI A, COSTANTINO R, WORTSMAN J, PAUS R, AND LING
related to corticotropin-releasing factor in the rat isolated basilar N. Melanotropic activity of gamma MSH peptides in melanoma
artery. Br J Pharmacol 125: 1164 –1171, 1998. cells. Life Sci 50: 1103–1108, 1992.
339. SCHIOTH HB, MUCENIECE R, LARSSON M, AND WIKBERG JE. 359. SLOMINSKI A, ERMAK G, HWANG J, CHAKRABORTY A, MA-
The melanocortin 1, 3, 4 or 5 receptors do not have a binding ZURKIEWICZ JE, AND MIHM M. Proopiomelanocortin, cortico-
epitope for ACTH beyond the sequence of alpha-MSH. J Endocri- tropin releasing hormone and corticotropin releasing hormone re-
nol 155: 73–78, 1997. ceptor genes are expressed in human skin. FEBS Lett 374: 113–116,
340. SCHIOTH HB, MUTULIS F, MUCENIECE R, PRUSIS P, AND WIK- 1995.
BERG JE. Selective properties of C- and N-terminals and core 360. SLOMINSKI A, ERMAK G, HWANG J, MAZURKIEWICZ JE, COR-
residues of the melanocyte-stimulating hormone on binding to the LISS D, AND EASTMAN A. The expression of proopiomelanocortin
human melanocortin receptor subtypes. Eur J Pharmacol 2–3: (POMC) and of corticotropin releasing hormone receptor (CRH-R)
359 –366, 1998. genes in mouse skin. Biochim Biophys Acta 1289: 247–251, 1996.
341. SCHOLZEN TE, BRZOSKA T, KALDEN DH, HARTMEYER M, 361. SLOMINSKI A, ERMAK G, MAZURKIEWICZ JE, BAKER J, AND
LUGER TA, ARMSTRONG CA, AND ANSEL JC. Expression of pro- WORTSMAN J. Characterization of corticotropin releasing hor-
opiomelanocortin peptides and prohormone convertases by human mone (CRH) in human skin. J Clin Endocrinol Metab 83: 1020 –
dermal microvascular endothelial cells. Ann NY Acad Sci 885: 1024, 1998.
444 – 447, 1999. 362. SLOMINSKI A, ERMAK G, AND MIHM M. ACTH receptor, CYP11A1,
342. SCHOLZEN TE, BRZOSKA T, KALDEN DH, HARTMEYER M, FAS- CYP17 and CYP21A2 genes are expressed in skin. J Clin Endocri-
TRICH M, LUGER TA, ARMSTRONG CA, AND ANSEL JC. Expres- nol Metab 81: 2746 –2749, 1996.
sion of functional melanocortin receptors and proopiomelanocor- 363. SLOMINSKI A, HEASLEY D, MAZURKIEWICZ JE, ERMAK G,
tin peptides by human dermal microvascular endothelial cells. Ann BAKER J, AND CARLSON JA. Expression of proopiomelanocortin
NY Acad Sci 885: 239 –253, 1999. (POMC) derived melanocyte stimulating hormone (MSH) and
343. SCOTT AP AND LOWRY PJ. Adrenocorticotrophic- and melanocyte- ACTH peptides in skin of basal cell carcinoma patients. Hum
stimulating peptides in the human pituitary. Biochem J 139: 593– Pathol 30: 208 –215, 1999.
602, 1974. 364. SLOMINSKI A, JASTREBOFF P, AND PAWELEK J. L-Tyrosine stim-
344. SEASHOLTZ A, BOURBONAIS F, HARNDEN C, AND CAMPER S. ulates induction of tyrosinase activity by MSH and reduces coop-
Nucleotide sequence and expression of the mouse corticotropin- erative interactions between MSH receptors in hamster melanoma
releasing hormone gene. Mol Cell Neurosci 2: 266 –273, 1991. cells. Biosci Rep 9: 579 –586, 1989.
345. SEIDAH NG, FOURNIER H, BOILEAU G, BENJANNET S, RON- 365. SLOMINSKI A AND MIHM M. Potential mechanism of skin response
DEAU N, AND CHRETIEN M. The cDNA structure of the porcine to stress. Int J Dermatol 35: 849 – 851, 1996.
pro-hormone convertase PC2 and the comparative processing by 366. SLOMINSKI A, MOELLMANN G, AND KUKLINSKA E. MSH inhibits
PC1 and PC2 of the N-terminal glycopeptide segment of porcine growth in a line of amelanotic hamster melanoma cells and induces
POMC. FEBS Lett 310: 235–239, 1992. increases in cAMP levels and tyrosinase activity without inducing
346. SHAHABI AN, BURTNESS MZ, AND SHARP BM. N-acetyl-␤-endor- melanogenesis. J Cell Sci 92: 551–559, 1989.
phin1–31 antagonized the suppressive effect of ␤-endorphin1–31 on 367. SLOMINSKI A AND PAUS R. Bomirski melanomas: a versatile and
murine splenocyte proliferation via a noloxone-resistant receptor. powerful model for pigment cell and melanoma research. Int J
Biochem Biophys Res Commun 175: 936 –942, 1991. Oncol 2: 21–228, 1993.
347. SHALTS E, FENG JY, FERIN M, AND WARDLAW SL. ␣-Melanocyte- 368. SLOMINSKI A AND PAUS R. Melanogenesis is coupled to murine
stimulating hormone antagonized the neuroendocrine effects of anagen: toward new concepts for the role of melanocytes and the
corticotropin-releasing factor and interleukin-1␤ in the primate. regulation of melanogenesis in hair growth. J Invest Dermatol 101:
Endocrinology 131: 132–138, 1992. 90S–97S, 1993.

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


1018 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

369. SLOMINSKI A, PAUS R, AND COSTANTINO R. Differential expres- nitric oxide synthase by alpha-MSH. Proc Natl Acad Sci USA 92:
sion and activity of melanogenesis-related proteins during induced 8016 – 8020, 1995.
hair growth in mice. J Invest Dermatol 96: 172–179, 1991. 391. STENZEL P, KESTERSON R, YEUNG W, CONE RD, RITTENBERG
370. SLOMINSKI A, PAUS R, AND MAZURKIEWICZ JE. Proopiomelano- MB, AND STENZEL-POORE MP. Identification of a novel murine
cortin expression and potential function during induced hair receptor for corticotropin-releasing hormone expressed in the
growth in C57BL6 mouse. Ann NY Acad Sci 642: 459 – 461, 1991. heart. Mol Endocrinol 9: 637– 645, 1995.
371. SLOMINSKI A, PAUS R, AND MAZURKIEWICZ JE. Proopiomelano- 392. STENZEL-POORE M, VALE WW, AND RIVIER C. Relationship be-
cortin expression in the skin during induced hair growth in mice. tween antigen-induced immune stimulation and activation of the
Experientia 48: 50 –54, 1992. hypothalamic-pituitary-adrenal axis in the rat. Endocrinology 132:
372. SLOMINSKI A, PAUS R, PLONKA P, HANDJINSKI B, MAURER M, 1313–1318, 1993.
CHAKRABORTY A, AND MIHM MC JR. Pharmacological disruption 393. STEPHANOU A, FITZHARRIS P, KNIGHT RA, AND LIGHTMAN SL.
of hair follicle pigmentation as a model for studying the melanocyte Characteristics and kinetics of proopiomelanocortin mRNA ex-
response to and recovery from cytotoxic damage in situ. J Invest pression by human leucocytes. Brain Behav Immun 5: 319 –327,
Dermatol 106: 1203–1211, 1996. 1991.
373. SLOMINSKI A, PAUS R, PLONKA P, MAURER M, CHAKRABORTY 394. STEPHANOU A, JESSOP DS, KNIGHT RA, AND LIGHTMAN SL.
A, PRUSKI D, AND LUKIEWICZ S. Melanogenesis during the ana- Corticotropin-releasing factor-like immunoreactivity and mRNA in
gen-catagen-telogen transformation of the murine hair cycle. J In- human leukocytes. Brain Behav Immun 4: 67–73, 1990.
vest Dermatol 102: 862– 869, 1994. 395. STEPHANOU A, OKOSI A, KNIGHT RA, CHOWDREY HS, AND
374. SLOMINSKI A, PAUS R, AND SCHADERDORF D. Melanocytes are LATCHMAN DS. C/EBP activates the human corticotropin-releas-
sensory and regulatory cells of epidermis. J Theor Biol 164: 103– ing hormone gene promoter. Mol Cell Endocrinol 134: 41–50, 1997.
120, 1993. 396. STEVENS MY, CHALLIS JR, AND LYE SJ. Corticotropin-releasing
375. SLOMINSKI A, PAUS R, AND WORTSMAN J. Can some mela- hormone receptor subtype 1 is significantly up-regulated at the
notropins modulate keratinocyte proliferation? J Invest Dermatol time of labor in the human myometrium. J Clin Endocrinol Metab
97: 747, 1991. 83: 4107– 4115, 1998.
376. SLOMINSKI A, PAUS R, AND WORTSMAN J. On the potential role of 397. STURM RA, BOX NF, AND RAMSAY M. Human pigmentation genet-
proopiomelanocortin in skin physiology and pathology. Mol Cell ics: the difference is only skin deep. Bioessays 20: 712–721, 1998.
Endocrinol 93: C1–C6, 1993. 398. SUDA T, TOZAWA F, DOBASHI I, HORIBA N, OHMORI N, YA-
377. SLOMINSKI A AND PAWELEK J. MSH binding in Bomirski amela- MAKADO M, YAMADA M, AND DEMURA H. Corticotropin-releasing
notic hamster melanoma cells is stimulated by L-tyrosine. Biosci hormone, proopiomelanocortin, and glucocorticoid receptor gene
Rep 7: 949 –954, 1987. expression in adrenocorticotropin-producing tumors in vitro.
378. SLOMINSKI A AND PAWELEK J. Animals under the sun: effects of J Clin Invest 92: 2790 –2795, 1993.
UV radiation on mammalian skin. Clin Dermatol 16: 503–515, 1998. 399. SUKHANOV VA, VORONKOVA IM, SHVETS SV, DYAKOV VL, AND
379. SLOMINSKI A, ROLOFF B, ZBYTEK B, WEI ET, FECHNER K, MOROZOVA LF. Melanocyte-stimulating hormone (␣-MSH) inhibits
CURRY J, AND WORTSMAN J. Corticotropin releasing hormone the growth of human malignant melanoma cells with the induction
(CRH) and related peptides can act as bioregulatory factors in of phosphatidylinositol and myo-inositol phosphate levels. Bio-
human keratinocytes. In Vitro Cell Dev Biol 36: 211–216, 2000. chem Int 24: 625– 632, 1991.
380. SLOMINSKI A, ROLOFF B, CURRY J, DAHIYA M, SZCZ- 400. SUZUKI I, CONE RD, IM S, NORDLUND J, AND ABDEL-MALEK ZA.
ESNIEWSKI A, AND WORTSMAN J. The skin produces urocortin. Binding of melanotropin hormones to the melanocortin receptor
J Clin Endocrinol Metab 85: 815– 823, 2000. MC1R on human melanocytes stimulates proliferation and melano-
381. SLOMINSKI A, WORTSMAN J, MAZURKIEWICZ JE, MATSUOKA genesis. Endocrinology 137: 1627–1633, 1996.
L, DIETRICH J, LAWRENCE K, GORBANI A, AND PAUS R. Detec- 401. SUZUKI I, TADA A, OLLMANN MM, BARSH GS, IM S, LAMOREUX
tion of the proopiomelanocortin-derived antigens in normal and ML, HEARING VJ, NORDLUND JJ, AND ABDEL-MALEK ZA. Agouti
pathologic human skin. J Lab Clin Med 122: 658 – 666, 1993. signaling protein inhibits melanogenesis and the response of hu-
382. SMITH A, BARCLAY C, QUABA A, SEDOWOFIA K, STEPHEN R, man melanocytes to alpha-melanotropin. J Invest Dermatol 108:
THOMPSON M, WATSON A, AND MCINTOSH N. The bigger the 838 – 842, 1997.
burn, the greater the stress. Burns 23: 291–294, 1997. 402. TAKAHASHI A, AMEMIYA Y, SARASHI M, SOWER SA, AND
383. SMITH AI AND FUNDER JW. Proopiomelanocortin processing in KAWAUCHI H. Melanotropin and corticotropin are encoded on two
the pituitary, central nervous system, and peripheral tissues. En- distinct genes in the lamprey, the earliest evolved extant verte-
docr Rev 9: 159 –179, 1988. brate. Biochem Biophy Res Commun 213: 490 – 498, 1995.
384. SMITH EM, HUGHES TK, HASHEMI F, AND STEFANO GB. Immu- 403. TAKAHASHI K, TOTSUNE K, SONE M, MURAKAMI O, SATOH F,
nosuppressive effects of corticotropin and melanotropin and their ARIHARA Z, SASANO H, IINO K, AND MOURI T. Regional distribu-
possible significance in human immunodeficiency virus infection. tion of urocortin-like immunoreactivity and expression of urocor-
Proc Natl Acad Sci USA 89: 782–786, 1992. tin mRNA in the human brain. Peptides 19: 643– 647, 1998.
385. SMITH EM, MORRILL AC, MEYER WJ, AND BLALOCK JE. Cortico- 404. TAKEUCHI S AND TAKAHASHI S. Melanocortin receptor genes in
tropin releasing factor induction of leukocyte-derived immunore- the chicken: tissue distributions. Gen Comp Endocrinol 112: 220 –
active ACTH and endorphins. Nature 321: 881– 882, 1986. 231, 1998.
386. SMITH R, HEALY E, SIDDIQUI S, FLANAGAN N, STEIJLEN PM, 405. TAMATE HB AND TAKEUCHI T. Action of the e locus of mice in the
ROSDAHL I, JACQUES JP, ROGERS S, TURNER R, JACKSON IJ, response of phaeomelanic hair follicles to ␣-melanocyte-stimulat-
BIRCH-MACHIN MA, AND REES JE. Melanocortin 1 receptor vari- ing hormone in vitro. Science 24: 1241–1242, 1984.
ants in an Irish population. J Invest Dermatol 111: 119 –122, 1998. 406. TATRO JB AND REICHLIN S. Specific receptors for ␣-melanocyte-
387. SMITH R, MESIANO S, CHAN E, BROWN S, AND JAFFE RB. stimulating hormone are widely distributed in tissues of rodents.
Corticotropin-releasing hormone directly and preferentially stimu- Endocrinology 121: 1900 –1907, 1987.
lates dehydroepiandrosterone sulfate secretion by human fetal ad- 407. TAYLOR AW, STREILEIN JW, AND COUSINS SW. Identification of
renal cortical cells. J Clin Endocrinol Metab 83: 2916 –2920, 1998. alpha-melanocyte stimulating hormone as a potential immunosup-
388. SPIESS J, DAUTZENBERG FM, SYDOW S, HAUGER RL, RUH- pressive factor in aqueous humor. Curr Eye Res 11: 1199 –1206,
MANN A, BLANK T, AND RADULOVIC J. Molecular properties of 1992.
the CRF receptor. Trends Endocrinol Metab 9: 140 –145, 1998. 408. TEOFOLI P, MOTOKI K, LOTTI TM, UITTO J, AND MAUVIEL A.
389. SPIRO J, PARKER S, OLIVER I, FRASER C, MARKS JM, AND Proopiomelanocortin (POMC) gene expression by normal skin and
THODY AJ. Effect of PUVA on plasma and skin immunoreactive keloid fibroblasts in culture: modulation by cytokines. Exp Derma-
alpha-melanocyte stimulating hormone concentrations. Br J Der- tol 6: 111–115, 1997.
matol 117: 703–707, 1987. 409. THEOHARIDES T, SINGH L, BOUCHER W, PANG X, LETOUR-
390. STAR RA, RAJORA N, HUANG J, STOCK RC, CATANIA A, AND NEAU R, WEBSTER E, AND CHROUSOS G. Corticotropin-releasing
LIPTON JM. Evidence of autocrine modulation of macrophage hormone induces skin mast cell degranulation and increased vas-

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


July 2000 CRH-POMC SIGNALING IN A CUTANEOUS STRESS RESPONSE 1019

cular permeability, a possible explanation for its proinflammatory of human corticotropin-releasing hormone gene expression: impli-
effects. Endocrinology 139: 403– 413, 1998. cations for the stress response and immune/inflammatory reaction.
410. THEOHARIDES T, SPANOS C, PANG X, ALFERES L, LIGRIS K, Endocr Rev 15: 409 – 420, 1994.
LETOURNEAU R, ROZNIECKI J, WEBSTER E, AND CHROUSOS G. 431. VAN DER KRAAN M, ADAN R, ENTWISTLE M, GISPEN W, BUR-
Stress-induced intracranial mast cell degranulation: a corticotrop- BACH JP, AND TATRO J. Expression of melanocortin-5 receptor in
in-releasing hormone-mediated effect. Endocrinology 136: 5745– secretory epithelia supports a functional role in exocrine and en-
5750, 1995. docrine glands. Endocrinology 139: 2348 –2355, 1998.
411. THODY AJ, RIDLEY K, PENNY RJ, CHALMERS R, FISHER C, AND 432. VAN TOL EA, PETRUZA P, LUND PK, YAMAUCHI M, AND SARTOR
SHUSTER S. MSH peptides are present in mammalian skin. Pep- RB. Local production of corticotropin releasing hormone is in-
tides 4: 813– 816, 1983. creased in experimental intestinal inflammation in rats. Gut 39:
412. THODY AJ AND SHUSTER S. Control and function of sebaceous 385–392, 1996.
glands. Physiol Rev 69: 383– 416, 1989. 433. VAN WIJK PA, VAN NECK JW, RIJNBERK A, CROUGHS RH, AND
413. THOMPSON RC, SEASHOLTZ AF, AND HERBERT E. Rat cortico- MOL HA. Proliferation of the murine corticotropic tumor cell line
tropin-releasing hormone gene: sequence and tissue-specific ex- AtT20 is affected by hypophysiotrophic hormones, growth factors
pression. Mol Endocrinol 1: 363–390, 1987. and glucocorticoids. Mol Cell Endocrinol 111: 13–19, 1995.
414. THORNTON JE, CHEUNG CC, CLIFTON DK, AND STEINER RA. 434. VARGA J, MOELLMANN G, FRITSCH P, GODAWSKA E, AND
Regulation of hypothalamic proopiomelanocortin mRNA by leptin LERNER A. Association of cell surface receptors for melanotropin
in ob/ob mice. Endocrinology 138: 5063–5066, 1997. with the Golgi region in mouse melanoma cells. Proc Natl Acad Sci
415. TIMPL P, SPANGEL R, SILLABER I, KRESSE A, REUL JM, STALLA USA 73: 559 –562, 1976.
GK, BLANQUET V, STECKLER T, HOLSBOER F, AND WURST W. 435. VARGA JM, DIPASQUALE A, PAWELEK J, MCGUIRE J, AND
Impaired stress response and reduced anxiety in mice lacking a LERNER A. Regulation of melanocyte-stimulating hormone (MSH)
functional corticotropin-releasing hormone receptor 1. Nature action at the receptor level: discontinuous binding of MSH to
Genet 19: 162–166, 1998. synchronized mouse melanoma cells during the cell cycle. Proc
416. TINAJERO JC, FABBRI A, AND DUFAU ML. Regulation of cortico- Natl Acad Sci USA 71: 1590 –1593, 1974.
tropin-releasing factor secretion from Leydig cells by serotonin. 436. VAUGHAN J, DONALDSON C, BITTENCOURT J, PERRIN MH,
Endocrinology 130: 1780 –1788, 1992. LEWIS K, SUTTON S, CHAN R, TURNBULL AV, LOVEJOY D,
417. TJUVAJEV J, KOLESNIKOV Y, JOSHI R, SHERINSKI J, RIVIER C, RIVIER J, SAWCHENKO PE, AND VALE W. Urocortin, a
KOUTCHER L, YZHOU, MATEI C, KOUTCHER J, KREEK MJ, AND mammalian neuropeptide related to fish urotensin I and to corti-
BLASBERG R. Anti-neoplastic properties of human corticotropin cotropin-releasing factor. Nature 378: 287–292, 1995.
releasing factor: involvement of the nitric oxide pathway. In Vivo 437. VITA N, LAURENT P, LEFORT S, CHALON P, LELIAS JM,
12: 1–10, 1998. KAGHAD M, LEFUR G, CAPUT D, AND FERRARA P. Primary
418. TORPY DJ AND CHROUSOS GP. The three-way interactions be- structure and functional expression of mouse pituitary and human
tween the hypothalamic-pituitary-adrenal and gonadal axes and the corticotropin releasing factors receptors. FEBS Lett 335: 1–5, 1993.
immune system. Baillieres Clin Rheumatol 10: 181–198, 1996. 438. WAKAMATSU K, GRAHAM A, COOK D, AND THODY JH. Charac-
419. TSAI-MORRIS CH, BUCZKO E, GENG Y, GAMBOA-PINTO A, AND terization of ACTH peptides in human skin and their activation of
DUFAU ML. The genomic structure of the rat corticotropin releas- melanocortin-1 receptor. Pigment Cell Res 10: 288 –297, 1997.
ing factor receptor. J Biol Chem 271: 14519 –14525, 1996. 439. WARDLAW SL, MCCARTHY KC, AND CONWELL IM. Glucocorticoid
420. TURNBULL A, SMITH G, LEE S, VALE WW, LEE K, AND RIVIER C. regulation of hypothalamic proopiomelanocortin. Neuroendocri-
CRF type I receptor-deficient mice exhibit a pronounced pituitary- nology 67: 51–57, 1998.
adrenal response to local inflammation. Endocrinology 140: 1013– 440. WEI ET, GAO GC, AND THOMAS HA. Peripheral anti-inflammatory
1017, 1999. actions of corticotropin-releasing factor. Ciba Found Symp 172:
421. TURNBULL AV AND RIVIER CL. Regulation of the hypothalamic- 258 –268, 1993.
pituitary-adrenal axis by cytokines: actions and mechanisms of 441. WEI ET AND THOMAS HA. Correlation of neuroendocrine and
action. Physiol Rev 79: 1–71, 1999. anti-edema activities of alanine-corticotropin-releasing factor ana-
422. TURNBULL AV, VALE W, AND RIVIER C. Urocortin, a corticotropin- logs. Eur J Pharmacol 263: 319 –321, 1994.
releasing factor-related mammalian peptide, inhibits edema due to 442. WEI ET, THOMAS HA, CHRISTIAN HC, BUCKINGHAM JC, AND
thermal injury in rats. Eur J Pharmacol 303: 213–216, 1996. ISHIMOTO TK. D-Amino acid-substituted analogs of corticotropin-
423. UEHARA Y, SHIMIZU H, SATO N, TANAKA Y, SHIMOMURA Y, AND releasing hormone (CRH) and urocortin with selective agonist
MORI M. Carboxyl-terminal tripeptide of ␣-melanocyte-stimulating activity at CRH1 and CRH2-beta receptors. Peptides 19: 1183–1190,
hormone antagonized interleukin-1-induced anorexia. Eur J Phar- 1998.
macol 220: 119 –122, 1992. 443. WESTLY HJ, KLEISS AJ, KELLEY KW, WONG PK, AND YUEN PH.
424. UHLER M AND HERBERT E. Complete amino acid sequence of Newcastle disease virus-infected splenocytes express the proopio-
mouse pro-opiomelanocortin derived from the nucleotide sequence melanocortin gene. J Exp Med 163: 1589 –1594, 1986.
of pro-opiomelanocortin cDNA. J Biol Chem 258: 257–261, 1983. 444. WILSON JD, FOSTER DW, KRONENBERG HM, AND LARSEN PR.
425. UHLER M, HERERT E, DEUSTACHIO P, AND RUDDLE FD. The Williams Textbook of Endocrinology (9th ed.). Philadelphia, PA:
mouse genome contains two nonallelic pro-opiomelanocortin Saunders, 1998.
genes. J Biol Chem 258: 9444 –9453, 1983. 445. WINTZEN M AND GILCHREST BA. Proopiomelanocortin, its de-
426. VALDENAIRE O, GILLER T, BREU V, GOTTOWIK J, AND KIL- rived peptides, and the skin. J Invest Dermatol 106: 3–10, 1996.
PATRICK G. A new functional isoform of the human CRF2 receptor 446. WINTZEN M, YAAR M, BURBACK JP, AND GILCHREST BA. Pro-
for corticotropin-releasing factor. Biochim Biophys Acta 1352: opiomelanocortin gene product regulation in keratinocytes. J In-
129 –132, 1997. vest Dermatol 106: 673– 678, 1996.
427. VALE W, SPIESS J, RIVER C, AND RIVER J. Characterization of a 447. WONG G AND PAWELEK J. Control of phenotypic expression of
41-residue ovine hypothalamic peptide that stimulates secretion of cultured melanoma cells by melanocyte-stimulating hormones. Na-
corticotropin and ␤-endorphin. Science 213: 1394 –1397, 1981. ture New Biol 241: 213–215, 1973.
428. VALVERDE P, HEALY E, JACKSON I, REES JL, AND THODY AJ. 448. WONG G AND PAWELEK J. Melanocyte-stimulating hormone acti-
Variants of the melanocyte-stimulating hormone receptor gene are vates preexisting tyrosinase molecules in mouse melanoma cells.
associated with red hair and fair skin in humans. Nature Genet 11: Nature 255: 644 – 646, 1975.
328 –330, 1995. 449. WONG G, PAWELEK J, SANSONE M, AND MOROWITZ J. Response
429. VALVERDE P, HEALY E, SIKKINK S, HALDANE F, THODY AJ, of mouse melanoma cells to melanocyte-stimulating hormone. Lo-
CAROTHERS A, JACKSON IJ, AND REES JL. The Asp84Glu variant calization in the G-2 phase of the cell cycle. Nature 248: 351–354,
of the melanocortin 1 receptor (MC1R) is associated with mela- 1974.
noma. Hum Mol Genet 5: 1663–1666, 1996. 450. YAWSEN L, DIEHL N, BRENNAN MB, AND HOCHGESCHWENDER
430. VAMVAKOPOULOS NC AND CHROUSOS GP. Hormonal regulation U. Obesity in the mouse model of pro-opiomelanocortin deficiency

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.


1020 SLOMINSKI, WORTSMAN, LUGER, PAUS, AND SOLOMON Volume 80

responds to peripheral melanocortin. Nature Med 5: 1066 –1070, tures of the mouse and human urocortin genes (Ucn and UCN).
1999. Genomics 50: 23–33, 1998.
451. YAMAZAKI T, KIMOTO T, HIGUCHI K, OHTA Y, KAWATO S, AND 455. ZHOU A, BLOOMQUIST BT, AND MAINS RE. The prohormone
KOMINAMI S. Calcium ion as a second messenger for o-nitrophe- convertases PC1 and PC2 mediate district endorproteolytic cleav-
nylsulfenyl-adrenocorticotropin (NPS-ACTH) and ACTH in bovine ages in a strict temporal order during proopiomelanocortin biosyn-
adrenal steroidogenesis. Endocrinology 139: 4765– 4771, 1998. thetic processing. J Biol Chem 268: 1763–1769, 1993.
452. YANG Y, THOMPSON DA, DICKINSON CJ, WILKEN J, BARSH GS, 456. ZOUMAKIS E, MARGIORIS AN, MAKRIGIANNAKIS A, STOUR-
KENT SB, AND GANTZ I. Characterization of agouti-related protein NARAS C, AND GRAVANIS A. Human endometrium as a neuroen-
binding to melanocortin receptors. Mol Endocrinol 13: 148 –155, 1999. docrine tissue: expression, regulation and biological roles of endo-
453. ZAGON IS, WU Y, AND MCLAUGHLIN PG. The opioid growth factor, metrial corticotropin-releasing hormone (CRH) and opioid
met-enkephalin, and the zeta opioid receptor are present in human peptides. J Endocrinol Invest 20: 158 –167, 1997.
and mouse skin and tonically act to inhibit DNA synthesis in the 457. ZUBAIR A, LAKSHMI MS, AND SHERBET GV. Expression of alpha-
epidermis. J Invest Dermatol 106: 490 – 497, 1996. melanocyte stimulating hormone and the invasive ability of the B16
454. ZHAO L, DONALDSON CJ, SMITH GW, AND VALE WW. The struc- murine melanoma. Anticancer Res 12: 399 – 402, 1992.

Downloaded from journals.physiology.org/journal/physrev (031.221.180.186) on July 11, 2023.

You might also like