banner



Tactile Receptors In The Skin

  • Periodical List
  • Cold Spring Harb Perspect Med
  • v.4(6); 2014 Jun
  • PMC4031957

Cold Spring Harb Perspect Med. 2014 Jun; 4(6): a013656.

Diversification and Specialization of Touch Receptors in Peel

David Thousand. Owens

aneDepartment of Dermatology, Columbia University Higher of Physicians and Surgeons, New York, New York 10032

2Department of Pathology and Cell Biology, Columbia University College of Physicians and Surgeons, New York, New York 10032

Ellen A. Lumpkin

1Department of Dermatology, Columbia University College of Physicians and Surgeons, New York, New York 10032

3Section of Physiology and Cellular Biophysics, Columbia University College of Physicians and Surgeons, New York, New York 10032

Abstruse

Our skin is the furthest outpost of the nervous organisation and a master sensor for harmful and innocuous external stimuli. As a multifunctional sensory organ, the skin manifests a diverse and highly specialized array of mechanosensitive neurons with complex terminals, or cease organs, which are able to discriminate different sensory stimuli and encode this information for advisable central processing. Historically, the basis for this diversity of sensory specializations has been poorly understood. In addition, the human relationship between cutaneous mechanosensory afferents and resident skin cells, including keratinocytes, Merkel cells, and Schwann cells, during the evolution and part of tactile receptors has been poorly defined. In this article, we will hash out conserved tactile end organs in the epidermis and hair follicles, with a focus on contempo advances in our understanding that have emerged from studies of mouse hairy skin.

Peel is our body's protective roofing and our largest sensory organ. Unique among our sensory systems, the skin's nervous arrangement gives rise to distinct sensations, including gentle touch, pain, itch, warmth, and cold. These distinct percepts are initiated past an impressive array of somatosensory neurons, whose sensory axons, called afferents, densely innervate the skin (Fig. i). We rely on sensory inputs from the peel to collaborate with objects in our environs and to avoid harm.

An external file that holds a picture, illustration, etc.  Object name is cshperspectmed-SKN-013656_F1.jpg

Touch receptors of hairy skin. A diverse group of mechanosensory afferents innervate the hairy skin of mammals. The schematic depicts anatomically singled-out finish organs (right), which requite rise to neural signals with distinct patterns of activity (heart). These different classes of sensory neurons initiate the perception of different cutaneous sensations (left). Aβ afferents (blue shades), which have thick myelin sheaths, are gentle-touch receptors that display apace adapting (RA) or slowly adapting (SA) responses to bear on. In hairy pare, RA afferents form lanecolate endings effectually hair follicles. Slowly adapting type I (SAI) afferents innervate Merkel cells (yellow) amassed in touch domes. Aδ afferents (green shades), which have thin myelin sheaths, include Aδ low-threshold mechanoreceptors (LTMRs) and A-mechanonociceptors (AM), whose morphological end organs have not been identified. C-afferents (magenta shades) include C-LTMRs that innervate hair follicles every bit well as pruritoceptors and nociceptors that innervate the epidermis. (Modified from Bautista and Lumpkin 2011.)

Our sense of touch enables us to perform numerous behaviors that rely on fine motor skills, including typing, feeding, and dressing ourselves. Impact is also important for social exchange, including pair bonding and child rearing (Tessier et al. 1998; Feldman et al. 2010). Infants deprived of touch stimuli display developmental and cognitive deficits (reviewed in Kaffman and Meaney 2007; Ardiel and Rankin 2010). For example, premature babies show delayed development and growth but this can exist improved by 45 minutes of daily impact stimulation. Cognitive deficits in bear upon-deprived rodent pups persist through adulthood, highlighting the importance of touch during development.

Acute itch and pain are warning signals. Pain alerts us to noxious mechanical, chemical, and thermal stimuli that have the potential to impairment skin tissue. Moreover, under conditions of inflammation or injury, skin displays a hypersensitivity to impact and temperature that encourages usa to protect injured areas. Although the evolutionary advantage of itch is not fully understood, this awareness might provide protection past alerting us to the presence of insects that have the potential to transmit disease. Itch sensation involves complex signaling betwixt keratinocytes, allowed cells, and sensory neurons that innervate the epidermis (Liu and Ji 2013; Wilson et al. 2013).

In chronic pain and itch, these unpleasant sensations persist across the threat of tissue injury. These chronic states back-trail numerous pathophysiological weather condition, leading to pain triggered past gentle bear upon (allodynia), enhanced sensitivity to noxious mechanical or thermal stimuli (hyperalgesia), and unrelieved itch (Gilron et al. 2006; Liu and Ji 2013). These are prevalent complaints in developed countries. Chronic pain is estimated to agonize more than xxx% of Americans (Johannes et al. 2010). Moreover, unrelieved itch is one of the most mutual reasons for dermatological consult (Summery 2009).

Among the skin's protective functions, our agreement of cutaneous sensations has lagged compared with barrier and immune functions. Contempo studies of mouse models have advanced our knowledge of the interactions between skin and the nervous system that drives sensation. Hither, we focus on two mammalian skin regions that are specialized for carrying affect stimuli to the nervous system—the touch dome and the hair follicle. For additional insights into mechanosensory transduction, nociception, and crawling, we refer the reader to excellent recent reviews (Basbaum et al. 2009; Chalfie 2009; Jeffry et al. 2011; Patel and Dong 2011; Liu and Ji 2013).

MAMMALIAN SKIN IS INNERVATED BY Singled-out TYPES OF SENSORY NEURONS

Cutaneous sensory afferents brandish a diversity of developmental programs, molecular receptors, anatomical specializations, and neural signals, which let them to trigger distinct sensations (Fig. one). At the molecular level, dissimilar classes of cutaneous sensory neurons are specified by combinatorial expression of transcription factors and distinct neurotrophin dependencies (reviewed in Liu and Ma 2011; Lallemend and Ernfors 2012). Moreover, they express different receptors that can respond selectively to chemicals, temperatures, or physical forces (Lumpkin and Caterina 2007; Rice and Albrecht 2008). When these sensory receptors are activated, they atomic number 82 to membrane potential changes that trigger discrete neural signals in the form of activeness potential trains with unique patterns of activity. These neural signals are then processed by spinal cord and encephalon circuitry to produce singled-out percepts.

Sensory afferents can be distinguished anatomically based on their sensory terminals, or end organs, in the skin (Fig. ane). Unmyelinated afferents have complimentary nervus endings that finish in protrusions between epidermal keratinocytes. These include pruritoceptors, which trigger itch; nociceptors, which evoke painful sensations; and thermoreceptors, which respond to pare heating or cooling. In addition to their afferent function, many of these neurons serve an efferent role to modulate peel cells under conditions of inflammation or injury. When excited by a painful or itch-producing stimulus, peptidergic sensory neurons can release signaling molecules, such as substance P, calcitonin factor-related peptide (CGRP), and inflammatory mediators, from their gratis nerve endings. Thus, the complex interplay of neurons and skin cells is an integral component of skin function.

Whereas nociceptors and pruritoceptors have free nervus endings, most tactile afferents have terminal specializations that shape mechanosensory responses and so that different features of a tactile stimulus are represented in their neural firing patterns (de Garavilla et al. 2001; Rice and Albrecht 2008; Gardner et al. 2013). For example, a surprising variety of rapidly adapting afferents innervate hair follicles to respond to brushing, stroking, or air movements (Li et al. 2011). SA afferents respond to steady pressure level by producing discharges throughout skin stimulation. The best characterized of these is the SA type I (SAI) afferent, which forms complexes with Merkel cells located in the stratum basale of the epidermis. Other types of SA afferents also innervate skin; however, their end organs take non been identified in mice. Although the correspondence between an end organ and its neural firing pattern has traditionally been correlative, recent studies have begun to identify molecular markers that can exist used to reliably distinguish afferent subtypes in mouse skin (Loewenstein and Rathkamp 1958; Woodbury and Koerber 2007; Bourane et al. 2009; Luo et al. 2009; Seal et al. 2009; Li et al. 2011; Abraira and Ginty 2013).

Cutaneous afferents can as well be classified based on the speed with which they conduct action potentials, which is set past their degree of myelination (Brown and Iggo 1967; Rice and Albrecht 2008; Gardner et al. 2013). Aβ afferents are rapidly conducting fibers that have large axonal diameters and thick myelin sheaths. C fibers, which have thin, unmyelinated afferents, are the slowest class. Aδ fibers, which have fine axonal diameters and thin myelin sheaths, display intermediate conduction velocities. Although there are many exceptions, most nociceptors, pruritoceptors, and thermoreceptors are classified every bit Aδ or C fibers. Nigh tactile afferents, or low-threshold mechanoreceptors (LTMRs), autumn into Aβ or Aδ categories. A notable exception is C-LTMRs, a class of unmyelinated, LTMRs that abundantly innervate hairy skin and reply to gentle brushing (Olausson et al. 2010; Vrontou et al. 2013). These afferents have been proposed to mediate social aspects of bear upon that contribute to pair and maternal bonding (Olausson et al. 2010).

MERKEL Prison cell–NEURITE COMPLEXES IN Touch on DOMES

Merkel cell–neurite complexes respond to pressure and encode an object's spatial features, such as shapes and edges (Johnson 2001). These complexes are located in peel areas specialized for high tactile vigil, including glabrous fingerpads, vibrissal follicles, and touch domes, which are high-sensitivity areas of hairy skin (Fig. 2).

An external file that holds a picture, illustration, etc.  Object name is cshperspectmed-SKN-013656_F2.jpg

Position and construction of affect domes in mice. (A) Overhead view schematic illustrates that impact domes (TDs) are asymmetrical crescent-shaped structures that are polarized to the caudal side of tylotrich (guard) hairs in the pelage skin. (B) Sagittal view schematic illustrating the cardinal cellular and structural elements of the touch dome including unusual columnar keratinocytes juxtaposed with mature Merkel cells, which are innervated past SAI sensory afferents.

In the hairy skin of rodents and humans, affect domes (TDs) are specialized structures in the epidermis that consist of unusual columnar keratinocytes juxtaposed to Merkel cells (Fig. 2 ) (Pinkus 1902; Smith 1977; Moll et al. 1996b; Halata et al. 2003; Reinisch and Tschachler 2005). TDs are asymmetric, crescent-shaped structures that are typically associated with tylotrich (guard) pilus follicles. In the developed mouse, Merkel cell clusters are polarized to the caudal side of tylotrich hairs and this polar organization has recently been shown to require Frizzled6 signaling (Chang and Nathans 2013). In human skin, TDs are most abundant on the trunk and about half are associated with pilus follicles (Orime et al. 2013). Histological studies propose that man TDs are innervated by multiple types of sensory afferents (Reinisch and Tschachler 2005). A recent study institute that human TDs range in diameter from ∼50 to 500 µm and incorporate 65–265 Merkel cells (Orime et al. 2013). In adult mice, TDs are somewhat smaller, containing ∼5–40 Merkel cells in a cluster ≤100 µm in bore (Lumpkin et al. 2003; Lesko et al. 2013). Although the molecular and cellular ground for the designation and patterning of bear on dome cells during skin evolution remains poorly understood, recent evidence shows that touch dome keratinocytes share like keratin markers with hair follicle keratinocytes (Doucet et al. 2013), indicating that TDs may exist designated in the hair placode during morphogenesis.

The function of Merkel cells in the epidermis has been debated for decades. A long-held model posits that Merkel cells are mechanosensory cells that transduce touch and actuate SAI afferents through synaptic transmission (Iggo and Muir 1969; Tachibana and Nawa 2002; Haeberle and Lumpkin 2008). Consistent with this model, several groups have shown that Merkel cells are activated by mechanical stimuli, such every bit cell swelling and membrane stretch (Chan et al. 1996; Haeberle et al. 2008; Boulais et al. 2009). Moreover, parallels are notable betwixt Merkel cells and other mechanosensory receptor cells, such as pilus cells of the inner ear. For example, these cell types express a common complement of transcription factors, including Brn3c, Gfi1, Sox2, and mammalian atonal homolog 1 (Atoh1) (Xiang et al. 1997; Ben-Arie et al. 2000; Leonard et al. 2002; Wallis et al. 2003; Haeberle et al. 2004; Badea et al. 2012; Lesko et al. 2013). Atoh1 is required for developmental specification of both pilus cells and Merkel cells (Bermingham et al. 1999; Maricich et al. 2009; Van Keymeulen et al. 2009). A 2nd hypothesis is that Merkel cells are accessory cells that serve every bit mechanical filters or release neuromodulators to shape the firing patterns of impact-sensitive SAI afferents (Gottschaldt and Vahle-Hinz 1981). A two-receptor site model, which posits that Merkel cells and sensory terminals contribute to dissimilar aspects of the SAI response, has also to be proposed (Yamashita and Ogawa 1991). Finally, the presence of noninnervated Merkel cells in some epithelia suggests that Merkel cells might participate in neuroendocrine rather than sensory functions (reviewed in Eispert et al. 2009).

Consequent with an active office in touch on reception, a substantial body of histological and molecular show indicates that Merkel cells make synaptic-like contacts with sensory afferents (Haeberle et al. 2004; Hitchcock et al. 2004; Nunzi et al. 2004). Moreover, Merkel cells have pocket-size dense-core vesicles that incorporate neurotransmitters, including glutamate, ATP, serotonin, and a multitude of neuropeptides (Hartschuh et al. 1979, 1983; Alvarez et al. 1988; Gauweiler et al. 1988; Hartschuh and Weihe 1988; Garcia-Caballero et al. 1989; Toyoshima and Shimamura 1991; English et al. 1992; Fagan and Cahusac 2001; Tachibana and Nawa 2002; Haeberle et al. 2004; Hitchcock et al. 2004). A directly demonstration that Merkel cells release these neurotransmitters to convey sensory information to SAI afferents is even so defective. Alternatively, Merkel cells could release these substances in a paracrine fashion to modulate the development or function of neurons, keratinocytes, or other skin cells.

The requirement for Merkel cells in bear on has been tested using transgenic mice and photoablation (Ogawa 1996; Halata et al. 2003). Attempts to remove Merkel cells via photoablation or enzymatic digestion have produced conflicting results, with loss of SA touch responses in some studies but not others (Diamond et al. 1988; Ikeda et al. 1994; Mills and Diamond 1995; Senok et al. 1996a,b). Because information technology is difficult to completely ablate Merkel cells without also damaging sensory terminals with these methods, other groups take turned to transgenic approaches. In mutant mice lacking the p75 neurotrophin receptor, TDs have normal complements of Merkel cells at birth only virtually are lost postnatally. Electrophysiological recordings showed that these mutants have affect-evoked responses similar to those of wild-type mice (Kinkelin et al. 1999), indicating that Merkel cells are non necessary to produce slowing adapting touch responses. One caveat is that the presence of TDs was not confirmed in the receptive fields of these afferents; therefore, it is possible that these responses reflected the activity of SA neurons other than SAI afferents (Wellnitz et al. 2010). By dissimilarity, in Atoh1 provisional knockout mice, Merkel cells fail to course during development but TDs are nonetheless innervated by sensory afferents (Maricich et al. 2009). Merkel cell knockout mice testify a selective loss of SAI responses in electrophysiological recordings and a loss of texture preference in behavioral assays (Maricich et al. 2009, 2012). Thus, these results propose that Merkel cells play an integral role in touch-evoked SAI responses (Maricich et al. 2009). Because Merkel cells never develop in these mice, these studies do non distinguish between a developmental requirement, a mechanosensory function, or an accessory role. Thus, further studies are needed to define the function of Merkel cells in bear upon reception.

Merkel cells limited numerous neuronal proteins (Halata et al. 2003; Haeberle et al. 2004), which is consequent with the notion that the developmental origin of the Merkel lineage is the neural crest. This idea was initially supported by cell-lineage tracing studies identifying Wnt1-expressing neural crest stem cells as the Merkel-cell site of origin (Szeder et al. 2003). More recently, two reports comparing murine lineage tracing models using a neural crest Cre driver (Wnt1Cre) or an epidermal Cre commuter (K14Cre) indicated that Merkel cells are derived from the proliferative keratinocyte layer (Krt14-expressing) of skin rather than Wnt1 progenitors in the neural crest (Morrison et al. 2009; Van Keymeulen et al. 2009). In support of these findings, conditional deletion of Atoh1 in K14Cre mice was shown to abolish Merkel jail cell development, whereas the same mutation using Wnt1Cre mice had no effect (Morrison et al. 2009).

Although we understand very little about the signaling pathways that designate a Merkel fate in keratinocyte stem cells, recent studies take begun to unravel transcription factors other than Atoh1 that are required for Merkel cell differentiation. Conditional deletion of Sox2, a marker of Merkel cells (Haeberle et al. 2004; Driskell et al. 2009; Lesko et al. 2013), via K14Cre mice leads to depletion of Merkel cells in the epidermis (Bardot et al. 2013; Lesko et al. 2013). Interestingly, Sox2 signaling appears to be upstream of Atoh1 and is suppressed by Ezh1 and Ezh2 histone methyltransferase enzymes (Bardot et al. 2013).

Considering Merkel cells are postmitotic cells (Vaigot et al. 1987; Merot and Saurat 1988; Moll et al. 1996c; Woo et al. 2010), an epidermal progenitor pool would presumably be required to maintain this lineage during epidermal homeostasis. Indeed, a phenotypically distinct population of columnar keratinocytes residing in TDs possesses bipotent progenitor capacity, equally evidenced by their ability to contribute to mature Merkel and squamous epidermal lineages during homeostasis and under regenerative conditions (Woo et al. 2010). Recently, a mouse model has been reported that targets touch dome columnar keratinocytes, while excluding mature Merkel cells, with a tamoxifen-inducible Cre recombinase (CreERT2) under the regulatory control of the cytokeratin Krt17 locus (Doucet et al. 2013). Lineage studies in Krt17CreERT2 transgenic mice showed that Merkel cells are genetic descendants of Krt17+ touch dome keratinocytes and that the puddle of Merkel cells in the impact dome is turned over every ii months in murine skin during homeostasis (Doucet et al. 2013). Interestingly, there does not appear to be whatsoever remarkable overlap between the touch dome niche and other niches in the remainder of the epidermis, indicating that the bear upon dome might correspond a distinct stalk cell puddle in the interfollicular epidermis (Doucet et al. 2013).

Our understanding of the cellular basis for the maintenance of Merkel prison cell homeostasis is highly relevant for two pathological skin conditions: Merkel prison cell carcinoma (MCC), which features an overproduction of neoplastic Merkel cells, and age-related loss of tactile acuity, which is associated with a loss of Merkel cells. MCC is a highly aggressive tumor that arises in the pare (Toker 1972) and consists of neoplastic cells expressing neuroendocrine markers and transcription factors like to those of normal Merkel cells (Harms et al. 2013). Although MCCs are relatively rare compared with other types of nonmelanoma skin cancer, the incidence of MCC cases has tripled over the last decade with less than one-half of MCC patients surviving v years following detection of lymph node interest (Hodgson 2005). MCC incidence is at least ten-fold greater in human immunodeficiency virus (HIV) and organ transplant patients as well every bit patients exposed to ultraviolet radiations (Lunder and Stern 1998; Penn and Showtime 1999; Engels et al. 2002), indicating that immune suppression may be a disquisitional risk factor for MCC. A previously unidentified polyoma virus, Merkel cell polyoma virus (MCV), was found in 80% of MCC cases (Feng et al. 2008) suggesting that opportunistic MCV infection plays a causal role in the formation of MCC; however, the pathogenesis of MCC remains poorly divers.

One issue clouding our understanding of MCC pathogenesis stems from conflicting reports addressing whether MCC actually arises from Merkel cells that have undergone oncogenic transformation (Gould et al. 1985; Kanitakis et al. 1998; Sidhu et al. 2005). Numerous neuronal proteins expressed by normal Merkel cells are likewise featured in MCC cells, supporting the idea of a Merkel cell origin for MCC. On the other hand, Merkel cells are postmitotic cells (Vaigot et al. 1987; Merot and Saurat 1988; Moll et al. 1996a) making them an unlikely candidate every bit a MCC cell of origin. In improver, the mature Merkel cell pool has a relatively fast (2 months) rate of turnover in adult murine peel (Doucet et al. 2013), which raises the possibility that an epidermal keratinocyte stem cell pool may be an origin for MCC.

INNERVATION OF HAIR FOLLICLES

Hair follicles are a major component of the pilosebaceous unit, which performs a number of functions in the pare including production of pilus fibers and oil, induction of angiogenesis, serving as a reservoir for pigment-producing and immune cells, and participation in wound healing. Recently, the hair follicle has gained more than attention for its role in touch sensation. The displacement of hairs by innocuous mechanical stimuli activates LTMRs housed in a piloneural collar that is located in the isthmus and upper bulge regions of the hair follicle (Figs. 1 and three). The terminals of inner piloneural afferent populations align with the cervix of the pilus follicle in a longitudinal direction forming palisades of lanceolate nerve endings, and outer populations innervate in a circumferential pattern (Munger and Ide 1988; Halata 1993).

An external file that holds a picture, illustration, etc.  Object name is cshperspectmed-SKN-013656_F3.jpg

Schematic representation of the structural components of the piloneural neckband mechanoreceptor. Sensory projections from dorsal root ganglion neurons innervate the isthmus/upper bulge region of hair follicles in the pelage skin in circumferential (CF) and longitudinal (LF) patterns. The terminal endings of these sensory afferents are tightly associated with the upward processes of terminal Schwann cells (tSCs; red). Both the fibers and tIISC processes are associated with the outer root sheath keratinocytes in the pilus follicle. The four types of pelage follicles, zigzag, awl, auchene, and baby-sit, and the accompanying afferent combinations are shown. Bg, bulge; CF, circumferential fibers; IFE, interfollicular epidermis; Is, isthmus; LF, lanceolate fibers; SG, sebaceous gland.

Whereas histological data indicate that at least five distinct populations of sensory neurons localize to the piloneural collar (Millard and Woolf 1988), recent studies using sparse genetic labeling have revealed that the repertoire of sensory receptors innervating mouse hairy skin is even more circuitous (Li et al. 2011; Badea et al. 2012; Wu et al. 2012; reviewed in Abraira and Ginty 2013). For example, at least x morphologically distinguishable types of sensory afferents innervate mouse dorsal skin at postnatal day 21 (Wu et al. 2012). Many of the sensory neurons that innervate pilus follicles express the transcription factor Brn3b and are derived from early Ret-positive sensory neurons during embroyonic evolution (Bourane et al. 2009; Luo et al. 2009; Badea et al. 2012). In some cases, an individual sensory afferent innervates only a single hair follicle, whereas other afferents innervate hundreds of hair follicles spanning a skin area of up to ∼7 mm2 (Li et al. 2011; Wu et al. 2012). The complicated composition of the piloneural collar might allow the hair follicle to discriminate between subtle differences in mechanical strength or the elapsing of stimulation.

Adding to this complication are the distinct pilus follicle types identified in murine skin—zigzag, awl, auchene, and guard—that tin can be discriminated based on hair shaft thickness, length, and the presence of kinks (Fig. three) (Schlake 2007). The discovery of selective molecular markers for unlike classes of hair follicle afferents has enabled a systematic analysis of the pilonural collars of these different hair follicle types (Li et al. 2011). Piloneural lanceolate endings are formed by at least three types of hair follicle sensory neurons: Aβ-LTMR, Aδ-LTMR, or C-LTMRs (Li et al. 2011). Each of these subtypes reports pilus movements but they convey information to the central nervous system with different latencies. Moreover, Aβ-LTMR, Aδ-LTMR, and C-LTMRs that innervate the same pare area display overlapping but singled-out central projection patterns, suggesting that their inputs are integrated by spinal cord circuitry. Remarkably, guard, awl/auchene, and zigzag follicles are innervated by unique complements of these sensory neurons in adolescent mice (P14–P30) (Fig. 3). These data suggest that each type of hair follicle represents a distinctive mechanosensory unit equipped to selectively encode specific aspects of brushing or stroking stimuli. These signals are then integrated by spinal circuits to form a cohesive representation of touch (Li et al. 2011; Abraira and Ginty 2013).

The palisade patterning of terminal nerve endings are a unique feature of the piloneural collar receptor, which appears to be influenced in part by the presence of type II terminal Schwann cells (tIISCs). A recent study reported that tIISCs are required for maintenance of mouse lanceolate endings and that these tIISCs persist post-obit denervation (Li and Ginty 2014). These tIISCs express Nestin and S100 and display long fingerlike processes that extend upwardly from tIISC jail cell bodies and interdigitate with longitudinal fibers so that each nerve ending is tightly juxtaposed on either side with tIISC processes (Kaidoh and Inoue 2000, 2008; Woo et al. 2010). Electron microscopic studies accept shown that Northward-cadherin-mediated adherens junctions are formed between outer root sheath (ORS) keratinocytes in the pilus follicle and either tIISC processes or the terminal nerve endings themselves (Kaidoh and Inoue 2000, 2008). These data bespeak that the maintenance of this receptor might rely on advice between all 3 cellular components. Support for this idea has come from analysis of the peel of mice lacking Vglut2, a vesicular glutamate transporter that packages excitatory glutamate into exocytic vesicles. Tissue-specific deletion of Vglut2 showed that neuron-derived glutamate plays an essential office in the evolution, maintenance, and mechanosensory capacity of the piloneural neckband (Woo et al. 2010). These studies illustrate that terminal Schwann cells might have a key role in the part of somatosensory receptors by facilitating the positioning of sensory stop organs. Collectively, these results showed that glutamate derived from sensory terminals is essential for the proper evolution, maintenance, and sensory function of the piloneural mechanoreceptor (Woo et al. 2012). This is the first evidence that excitatory glutamate derived from sensory neurons can regulate the differentiation of glial cells in the skin. Importantly, these results confer an efferent functionality to this population of glutamatergic sensory afferents in the pare. Glutamate release from exocytic vesicles has been observed following stretch activation of mechanosensory nerve terminals innervating musculus spindles, providing a precedence for this concept in other peripheral tissues. Moreover, sensory neuron-derived Shh ligands have been shown to modulate the ability of hair follicle stem cells to respond to skin wounding (Brownell et al. 2011). Whether additional efferent factors released by sensory neurons influence pare or hair follicle homeostasis or pathological skin atmospheric condition including cancer carcinogenesis remains poorly understood.

Final REMARKS

We are witnessing an exciting period of rapid progress in the field of cutaneous neurobiology. The combination of modern mouse genetics, neurophysiology, development, and stem prison cell biology has shed new light on the complex interactions betwixt the nervous system and skin cells, as well every bit the intricate neuronal processes that underlie our rich sensory experiences.

As highlighted in a higher place, fascinating questions remain unanswered. What are the signaling mechanisms in keratinocytes and other skin cell types that transduce efferent signals from cutaneous sensory neurons? What are the developmental pathways that let keratinocyte-derived Merkel cells to adopt a neuronal-like cell fate? Exercise epidermal Merkel cells and keratinocytes play a functional function in sensory signaling, equally suggested by numerous anatomical and molecular studies? Moreover, lilliputian is known nearly how the nervous system adapts to the changes in skin structure and function that accompanies normal aging and environmental exposures. For case, given that follicles tin produce different pilus types in adult hair cycles (Chi et al. 2013), do corresponding changes in innervation occur when pilus morphology switches? With modern tools in paw, the answers to these questions are at present within reach.

ACKNOWLEDGMENTS

The authors are supported by NIAMS R01 AR051219 (to E.A.L.), NINDS R01 NS073119 (to East.A.L. and Gregory J. Gerling), and NIEHS R21 ES020060 (to D.Thousand.O.). We give thanks Dr. Aislyn M. Nelson, Yanne Doucet, and Kara Mashall for assist with manuscript preparation. The authors declare no conflicts of interest.

Footnotes

REFERENCES

  • Abraira VE, Ginty DD 2013. The sensory neurons of bear upon. Neuron 79: 618–639 [PMC gratis article] [PubMed] [Google Scholar]
  • Alvarez FJ, Cervantes C, Blasco I, Villalba R, Martinez-Murillo R, Polak JM, Rodrigo J 1988. Presence of calcitonin factor-related peptide (CGRP) and substance P (SP) immunoreactivity in intraepidermal gratis nerve endings of cat peel. Brain Res 442: 391–395 [PubMed] [Google Scholar]
  • Ardiel EL, Rankin CH 2010. The importance of touch in development. Paediatr Kid Wellness 15: 153–156 [PMC free commodity] [PubMed] [Google Scholar]
  • Badea TC, Williams J, Smallwood P, Shi K, Motajo O, Nathans J 2012. Combinatorial expression of Brn3 transcription factors in somatosensory neurons: Genetic and morphologic analysis. J Neurosci 32: 995–1007 [PMC free article] [PubMed] [Google Scholar]
  • Bardot ES, Valdes VJ, Zhang J, Perdigoto CN, Nicolis S, Hearn SA, Silva JM, Ezhkova Due east 2013. Polycomb subunits Ezh1 and Ezh2 regulate the Merkel cell differentiation plan in skin stalk cells. EMBO J 32: 1990–2000 [PMC free article] [PubMed] [Google Scholar]
  • Basbaum AI, Bautista DM, Scherrer G, Julius D 2009. Cellular and molecular mechanisms of pain. Prison cell 139: 267–284 [PMC gratis commodity] [PubMed] [Google Scholar]
  • Bautista DM, Lumpkin EA 2011. Perspectives on: Data and coding in mammalian sensory physiology: Probing mammalian touch transduction. J Gen Physiol 138: 291–301 [PMC free commodity] [PubMed] [Google Scholar]
  • Ben-Arie N, Hassan BA, Bermingham NA, Malicki DM, Armstrong D, Matzuk Thousand, Bellen HJ, Zoghbi HY 2000. Functional conservation of atonal and Math1 in the CNS and PNS. Development 127: 1039–1048 [PubMed] [Google Scholar]
  • Bermingham NA, Hassan BA, Toll SD, Vollrath MA, Ben-Arie North, Eatock RA, Bellen HJ, Lysakowski A, Zoghbi HY 1999. Math1: An essential factor for the generation of inner ear hair cells. Science 284: 1837–1841 [PubMed] [Google Scholar]
  • Boulais N, Pennec JP, Lebonvallet Northward, Pereira U, Rougier N, Dorange G, Chesne C, Misery Fifty 2009. Rat Merkel cells are mechanoreceptors and osmoreceptors. PLoS ONE iv: e7759. [PMC complimentary article] [PubMed] [Google Scholar]
  • Bourane S, Garces A, Venteo S, Pattyn A, Hubert T, Fichard A, Puech Due south, Boukhaddaoui H, Baudet C, Takahashi S, et al. 2009. Low-threshold mechanoreceptor subtypes selectively express MafA and are specified by Ret signaling. Neuron 64: 857–870 [PubMed] [Google Scholar]
  • Dark-brown AG, Iggo A 1967. A quantitative report of cutaneous receptors and afferent fibres in the cat and rabbit. J Physiol 193: 707–733 [PMC gratuitous article] [PubMed] [Google Scholar]
  • Brownell I, Guevara East, Bai CB, Loomis CA, Joyner AL 2011. Nerve-derived sonic hedgehog defines a niche for hair follicle stem cells capable of becoming epidermal stalk cells. Cell Stem Cell 8: 552–565 [PMC free commodity] [PubMed] [Google Scholar]
  • Chalfie Yard 2009. Neurosensory mechanotransduction. Nat Rev Mol Jail cell Biol ten: 44–52 [PubMed] [Google Scholar]
  • Chan Due east, Yung WH, Baumann KI 1996. Cytoplasmic Ca2+ concentrations in intact Merkel cells of an isolated, functioning rat sinus hair preparation. Exp Brain Res 108: 357–366 [PubMed] [Google Scholar]
  • Chang H, Nathans J 2013. Responses of hair follicle-associated structures to loss of planar cell polarity signaling. Proc Natl Acad Sci 110: E908–E917 [PMC gratis commodity] [PubMed] [Google Scholar]
  • Chi W, Wu Eastward, Morgan BA 2013. Dermal papilla cell number specifies pilus size, shape and cycling and its reduction causes follicular decline. Evolution 140: 1676–1683 [PMC costless commodity] [PubMed] [Google Scholar]
  • de Garavilla L, Vergnolle N, Young SH, Ennes H, Steinhoff M, Ossovskaya VS, D'Andrea MR, Mayer EA, Wallace JL, Hollenberg Medico, et al. 2001. Agonists of proteinase-activated receptor 1 induce plasma extravasation past a neurogenic mechanism. Br J Pharmacol 133: 975–987 [PMC free article] [PubMed] [Google Scholar]
  • Diamond J, Mills LR, Mearow KM 1988. Evidence that the Merkel cell is not the transducer in the mechanosensory Merkel jail cell-neurite complex. Prog Brain Res 74: 51–56 [PubMed] [Google Scholar]
  • Doucet YS, Woo SH, Ruiz ME, Owens DM 2013. The affect dome defines an epidermal niche specialized for mechanosensory signaling. Cell Rep three: 1759–1765 [PMC gratuitous article] [PubMed] [Google Scholar]
  • Driskell RR, Giangreco A, Jensen KB, Mulder KW, Watt FM 2009. Sox2-positive dermal papilla cells specify hair follicle blazon in mammalian epidermis. Development 136: 2815–2823 [PMC free article] [PubMed] [Google Scholar]
  • Eispert AC, Fuchs F, Brandner JM, Houdek P, Wladykowski E, Moll I 2009. Evidence for distinct populations of human Merkel cells. Histochem Jail cell Biol 132: 83–93 [PubMed] [Google Scholar]
  • Engels EA, Frisch M, Goedert JJ, Biggar RJ, Miller RW 2002. Merkel prison cell carcinoma and HIV infection. Lancet 359: 497–498 [PubMed] [Google Scholar]
  • English KB, Wang ZZ, Stayner North, Stensaas LJ, Martin H, Tuckett RP 1992. Serotonin-like immunoreactivity in Merkel cells and their afferent neurons in touch domes from the hairy peel of rats. Anat Rec 232: 112–120 [PubMed] [Google Scholar]
  • Fagan BM, Cahusac PM 2001. Evidence for glutamate receptor mediated manual at mechanoreceptors in the peel. Neuroreport 12: 341–347 [PubMed] [Google Scholar]
  • Feldman R, Vocalizer 1000, Zagoory O 2010. Touch attenuates infants' physiological reactivity to stress. Dev Sci 13: 271–278 [PubMed] [Google Scholar]
  • Feng H, Shuda Yard, Chang Y, Moore PS 2008. Clonal integration of a polyomavirus in human Merkel cell carcinoma. Science 319: 1096–1100 [PMC gratuitous article] [PubMed] [Google Scholar]
  • Garcia-Caballero T, Gallego R, Roson E, Basanta D, Morel G, Beiras A 1989. Localization of serotonin-like immunoreactivity in the Merkel cells of grunter snout skin. Anat Rec 225: 267–271 [PubMed] [Google Scholar]
  • Gardner EP, Johnson KO 2013. Bear upon. In Principles of neural science (ed. Kandel ER, Schwartz JH, Jessell TM, Siegelbaum SA, Hudspeth AJ), pp. 498–529 McGraw-Hill, New York [Google Scholar]
  • Gauweiler B, Weihe E, Hartschuh Westward, Yanaihara N 1988. Presence and coexistence of chromogranin A and multiple neuropeptides in Merkel cells of mammalian oral mucosa. Neurosci Lett 89: 121–126 [PubMed] [Google Scholar]
  • Gilron I, Watson CP, Cahill CM, Moulin DE 2006. Neuropathic pain: A applied guide for the clinician. Can Med Assoc J 175: 265–275 [PMC complimentary article] [PubMed] [Google Scholar]
  • Gottschaldt KM, Vahle-Hinz C 1981. Merkel prison cell receptors: Structure and transducer function. Science 214: 183–186 [PubMed] [Google Scholar]
  • Gould VE, Moll R, Moll I, Lee I, Franke WW 1985. Neuroendocrine (Merkel) cells of the skin: Hyperplasias, dysplasias, and neoplasms. Lab Invest 52: 334–353 [PubMed] [Google Scholar]
  • Haeberle H, Lumpkin EA 2008. Merkel cells in somatosensation. Chem Percept 1: 110–118 [PMC free commodity] [PubMed] [Google Scholar]
  • Haeberle H, Fujiwara G, Chuang J, Medina MM, Panditrao M, Bechstedt S, Howard J, Lumpkin EA 2004. Molecular profiling reveals synaptic release machinery in Merkel cells. Proc Natl Acad Sci 101: 14503–14508 [PMC free article] [PubMed] [Google Scholar]
  • Haeberle H, Bryan LA, Vadakkan TJ, Dickinson ME, Lumpkin EA 2008. Swelling-activated Ca2+ channels trigger Ca2+ signals in Merkel cells. PLoS 1 three: e1750. [PMC free article] [PubMed] [Google Scholar]
  • Halata Z 1993. Sensory innervation of the hairy skin (light- and electronmicroscopic written report. J Invest Dermatol 101: 75S–81S [PubMed] [Google Scholar]
  • Halata Z, Grim M, Bauman KI 2003. Friedrich Sigmund Merkel and his "Merkel cell," morphology, development, and physiology: Review and new results. Anat Rec 271A: 225–239 [PubMed] [Google Scholar]
  • Harms Pow, Patel RM, Verhaegen ME, Giordano TJ, Nash KT, Johnson CN, Daignault S, Thomas DG, Gudjonsson JE, Elderberry JT, et al. 2013. Singled-out factor expression profiles of viral- and nonviral-associated Merkel jail cell carcinoma revealed past transcriptome assay. J Invest Dermatol 133: 936–945 [PMC complimentary commodity] [PubMed] [Google Scholar]
  • Hartschuh Due west, Weihe Due east 1988. Multiple messenger candidates and marker substance in the mammalian Merkel jail cell-axon circuitous: A light and electron microscopic immunohistochemical study. Prog Brain Res 74: 181–187 [PubMed] [Google Scholar]
  • Hartschuh W, Weihe Eastward, Buchler M, Helmstaedter V, Feurle GE, Forssmann WG 1979. Met enkephalin-like immunoreactivity in Merkel cells. Jail cell Tissue Res 201: 343–348 [PubMed] [Google Scholar]
  • Hartschuh Westward, Weihe E, Yanaihara North, Reinecke M 1983. Immunohistochemical localization of vasoactive intestinal polypeptide (VIP) in Merkel cells of diverse mammals: Evidence for a neuromodulator part of the Merkel jail cell. J Invest Dermatol 81: 361–364 [PubMed] [Google Scholar]
  • Hitchcock IS, Genever PG, Cahusac PM 2004. Essential components for a glutamatergic synapse between Merkel cell and nerve concluding in rats. Neurosci Lett 362: 196–199 [PubMed] [Google Scholar]
  • Hodgson NC 2005. Merkel cell carcinoma: Changing incidence trends. J Surg Oncol 89: 1–4 [PubMed] [Google Scholar]
  • Iggo A, Muir AR 1969. The structure and role of a slowly adapting impact corpuscle in hairy skin. J Physiol 200: 763–796 [PMC free article] [PubMed] [Google Scholar]
  • Ikeda I, Yamashita Y, Ono T, Ogawa H 1994. Selective phototoxic destruction of rat Merkel cells abolishes responses of slowly adapting type I mechanoreceptor units. J Physiol 479: 247–256 [PMC gratis commodity] [PubMed] [Google Scholar]
  • Jeffry J, Kim South, Chen ZF 2011. Crawling signaling in the nervous system. Physiology 26: 286–292 [PMC gratuitous article] [PubMed] [Google Scholar]
  • Johannes CB, Le TK, Zhou X, Johnston JA, Dworkin RH 2010. The prevalence of chronic pain in Us adults: Results of an Net-based survey. J Pain eleven: 1230–1239 [PubMed] [Google Scholar]
  • Johnson KO 2001. The roles and functions of cutaneous mechanoreceptors. Curr Opin Neurobiol 11: 455–461 [PubMed] [Google Scholar]
  • Johnson KO, Yoshioka T, Vega-Bermudez F 2000. Tactile functions of mechanoreceptive afferents innervating the hand. J Clin Neurophysiol 17: 539–558 [PubMed] [Google Scholar]
  • Kaffman A, Meaney MJ 2007. Neurodevelopmental sequelae of postnatal maternal intendance in rodents: Clinical and research implications of molecular insights. J Child Psychol Psychiatry 48: 224–244 [PubMed] [Google Scholar]
  • Kaidoh T, Inoue T 2000. Intercellular junctions betwixt palisade nerve endings and outer root sheath cells of rat vellus hairs. J Comp Neurol 420: 419–427 [PubMed] [Google Scholar]
  • Kaidoh T, Inoue T 2008. N-cadherin expression in palisade nervus endings of rat vellus hairs. J Comp Neurol 506: 525–534 [PubMed] [Google Scholar]
  • Kanitakis J, Bourchany D, Faure G, Claudy A 1998. Merkel cells in hyperplastic and neoplastic lesions of the skin. An immunohistochemical study using an antibody to keratin 20. Dermatology 196: 208–212 [PubMed] [Google Scholar]
  • Kinkelin I, Stucky CL, Koltzenburg M 1999. Postnatal loss of Merkel cells, but non of slowly adapting mechanoreceptors in mice defective the neurotrophin receptor p75. Eur J Neurosci 11: 3963–3969 [PubMed] [Google Scholar]
  • Lallemend F, Ernfors P 2012. Molecular interactions underlying the specification of sensory neurons. Trends Neurosci 35: 373–381 [PubMed] [Google Scholar]
  • Leonard JH, Cook AL, Van Gele G, Boyle GM, Inglis KJ, Speleman F, Sturm RA 2002. Proneural and proneuroendocrine transcription cistron expression in cutaneous mechanoreceptor (Merkel) cells and Merkel cell carcinoma. Int J Cancer 101: 103–110 [PubMed] [Google Scholar]
  • Lesko MH, Driskell RR, Kretzschmar K, Goldie SJ, Watt FM 2013. Sox2 modulates the function of two distinct cell lineages in mouse pare. Dev Biol 382: 15–26 [PMC gratis article] [PubMed] [Google Scholar]
  • Li L, Ginty DD 2014. The structure and system of lanceolate mechanosensory complexes at mouse hair follicles. eLife three: e01901. [PMC costless article] [PubMed] [Google Scholar]
  • Li 50, Rutlin M, Abraira VE, Cassidy C, Kus L, Gong S, Jankowski MP, Luo W, Heintz North, Koerber HR, et al. 2011. The functional arrangement of cutaneous low-threshold mechanosensory neurons. Cell 147: 1615–1627 [PMC costless article] [PubMed] [Google Scholar]
  • Liu T, Ji RR 2013. New insights into the mechanisms of itch: Are pain and itch controlled by distinct mechanisms? Pflugers Arch 465: 1671–1685 [PMC costless article] [PubMed] [Google Scholar]
  • Liu Y, Ma Q 2011. Generation of somatic sensory neuron diversity and implications on sensory coding. Curr Opin Neurobiol 21: 52–lx [PMC free article] [PubMed] [Google Scholar]
  • Loewenstein WR, Rathkamp R 1958. The sites for mechano-electrical conversion in a Pacinian corpuscle. J Gen Physiol 41: 1245–1265 [PMC free article] [PubMed] [Google Scholar]
  • Lumpkin EA, Caterina MJ 2007. Mechanisms of sensory transduction in the skin. Nature 445: 858–865 [PubMed] [Google Scholar]
  • Lumpkin EA, Collisson T, Parab P, Omer-Abdalla A, Haeberle H, Chen P, Doetzlhofer A, White P, Groves A, Segil N, et al. 2003. Math1-driven GFP expression in the developing nervous system of transgenic mice. Gene Expr Patterns 3: 389–395 [PubMed] [Google Scholar]
  • Lunder EJ, Stern RS 1998. Merkel-cell carcinomas in patients treated with methoxsalen and ultraviolet A radiations. N Engl J Med 339: 1247–1248 [PubMed] [Google Scholar]
  • Luo Westward, Enomoto H, Rice FL, Milbrandt J, Ginty DD 2009. Molecular identification of rapidly adapting mechanoreceptors and their developmental dependence on ret signaling. Neuron 64: 841–856 [PMC complimentary commodity] [PubMed] [Google Scholar]
  • Maricich SM, Wellnitz SA, Nelson AM, Lesniak DR, Gerling GJ, Lumpkin EA, Zoghbi HY 2009. Merkel cells are essential for calorie-free-touch responses. Science 324: 1580–1582 [PMC free article] [PubMed] [Google Scholar]
  • Maricich SM, Morrison KM, Mathes EL, Brewer BM 2012. Rodents rely on Merkel cells for texture bigotry tasks. J Neurosci 32: 3296–3300 [PMC free article] [PubMed] [Google Scholar]
  • Merot Y, Saurat JH 1988. Proliferation of Merkel cells in the pare. Acta Derm Venereol 68: 366–367 [PubMed] [Google Scholar]
  • Millard CL, Woolf CJ 1988. Sensory innervation of the hairs of the rat hindlimb: A light microscopic analysis. J Comp Neurol 277: 183–194 [PubMed] [Google Scholar]
  • Mills LR, Diamond J 1995. Merkel cells are non the mechanosensory transducers in the touch dome of the rat. J Neurocytol 24: 117–134 [PubMed] [Google Scholar]
  • Moll I, Gillardon F, Waltering S, Schmelz One thousand, Moll R 1996a. Differences of bcl-2 protein expression between Merkel cells and Merkel cell carcinomas. J Cutan Pathol 23: 109–117 [PubMed] [Google Scholar]
  • Moll I, Paus R, Moll R 1996b. Merkel cells in mouse peel: Intermediate filament pattern, localization, and hair cycle-dependent density. J Invest Dermatol 106: 281–286 [PubMed] [Google Scholar]
  • Moll I, Zieger W, Schmelz 1000 1996c. Proliferative Merkel cells were not detected in human being skin. Curvation Dermatol Res 288: 184–187 [PubMed] [Google Scholar]
  • Morrison KM, Miesegaes GR, Lumpkin EA, Maricich SM 2009. Mammalian Merkel cells are descended from the epidermal lineage. Dev Biol 336: 76–83 [PMC free article] [PubMed] [Google Scholar]
  • Munger BL, Ide C 1988. The structure and function of cutaneous sensory receptors. Curvation Histol Cytol 51: 1–34 [PubMed] [Google Scholar]
  • Nunzi MG, Pisarek A, Mugnaini E 2004. Merkel cells, corpuscular nerve endings and gratis nerve endings in the mouse palatine mucosa express iii subtypes of vesicular glutamate transporters. J Neurocytol 33: 359–376 [PubMed] [Google Scholar]
  • Ogawa H 1996. The Merkel prison cell equally a possible mechanoreceptor prison cell. Prog Neurobiol 49: 317–334 [PubMed] [Google Scholar]
  • Olausson H, Wessberg J, Morrison I, McGlone F, Vallbo A 2010. The neurophysiology of unmyelinated tactile afferents. Neurosci Biobehav Rev 34: 185–191 [PubMed] [Google Scholar]
  • Orime Thousand, Ushiki T, Koga D, Ito M 2013. Three-dimensional morphology of bear upon domes in human hairy skin by correlative calorie-free and scanning electron microscopy. J Invest Dermatol 133: 2108–2111 [PubMed] [Google Scholar]
  • Patel KN, Dong X 2011. Itch: Cells, molecules, and circuits. ACS Chem Neurosci 2: 17–25 [PMC free article] [PubMed] [Google Scholar]
  • Penn I, First MR 1999. Merkel'southward cell carcinoma in organ recipients: Written report of 41 cases. Transplantation 68: 1717–1721 [PubMed] [Google Scholar]
  • Pinkus F 1902. Uber einen bisher unbekannten Nebenapparat am Haarsystem des Menschen: Haarscheiben. Dermatologische Zeitschrift 9: 465–499 [Google Scholar]
  • Reinisch CM, Tschachler Eastward 2005. The touch dome in human skin is supplied by dissimilar types of nervus fibers. Ann Neurol 58: 88–95 [PubMed] [Google Scholar]
  • Rice FL, Albrecht PJ 2008. Cutaneous mechanisms of tactile perception: Morphological and chemical organization of the innervation to the pare. In The senses: A comprehensive reference (ed. Smith D, Firestein Due south, Beauchamp Grand). Academic, San Diego [Google Scholar]
  • Schlake T 2007. Determination of pilus structure and shape. Semin Cell Dev Biol 18: 267–273 [PubMed] [Google Scholar]
  • Seal RP, Wang X, Guan Y, Raja SN, Woodbury CJ, Basbaum AI, Edwards RH 2009. Injury-induced mechanical hypersensitivity requires C-low threshold mechanoreceptors. Nature 462: 651–655 [PMC free article] [PubMed] [Google Scholar]
  • Senok SS, Baumann KI, Halata Z 1996a. Selective phototoxic destruction of quinacrine-loaded Merkel cells is neither selective nor complete. Exp Brain Res 110: 325–334 [PubMed] [Google Scholar]
  • Senok SS, Halata Z, Baumann KI 1996b. Chloroquine specifically impairs Merkel cell mechanoreceptor function in isolated rat sinus hairs. Neurosci Lett 214: 167–170 [PubMed] [Google Scholar]
  • Sidhu GS, Chandra P, Cassai ND 2005. Merkel cells, normal and neoplastic: An update. Ultrastruct Pathol 29: 287–294 [PubMed] [Google Scholar]
  • Smith KR Jr 1977. The Haarscheibe. J Invest Dermatol 69: 68–74 [PubMed] [Google Scholar]
  • Summey BT 2009. Pruritis. In Pallative medicine (ed. Walsh D, Caraceni AT, Fainsinger R, Foley K, Glare P, Goh C, Lloyd-Williams M, Nunez Olarte J, Radbruch Fifty). Saunders, Philadelphia [Google Scholar]
  • Szeder V, Grim M, Halata Z, Sieber-Blum M 2003. Neural crest origin of mammalian Merkel cells. Dev Biol 253: 258–263 [PubMed] [Google Scholar]
  • Tachibana T, Nawa T 2002. Recent progress in studies on Merkel prison cell biological science. Anat Sci Int 77: 26–33 [PubMed] [Google Scholar]
  • Tessier R, Cristo M, Velez S, Giron M, de Calume ZF, Ruiz-Palaez JG, Charpak Y, Charpak Due north 1998. Kangaroo female parent care and the bonding hypothesis. Pediatrics 102: e17. [PubMed] [Google Scholar]
  • Toker C 1972. Trabecular carcinoma of the skin. Curvation Dermatol 105: 107–110 [PubMed] [Google Scholar]
  • Toyoshima G, Shimamura A 1991. Uranaffin reaction of Merkel corpuscles in the lingual mucosa of the finch, Lonchula striata var. domestica. J Anat 179: 197–201 [PMC free article] [PubMed] [Google Scholar]
  • Vaigot P, Pisani A, Darmon YM, Ortonne JP 1987. The majority of epidermal Merkel cells are not-proliferative: A quantitative immunofluorescence analysis. Acta Derm Venereol 67: 517–520 [PubMed] [Google Scholar]
  • Van Keymeulen A, Mascre Yard, Youseff KK, Harel I, Michaux C, De Geest N, Szpalski C, Achouri Y, Bloch W, Hassan BA, et al. 2009. Epidermal progenitors give ascension to Merkel cells during embryonic evolution and adult homeostasis. J Cell Biol 187: 91–100 [PMC free article] [PubMed] [Google Scholar]
  • Vrontou S, Wong AM, Rau KK, Koerber HR, Anderson DJ 2013. Genetic identification of C fibres that observe massage-like stroking of hairy skin in vivo. Nature 493: 669–673 [PMC complimentary article] [PubMed] [Google Scholar]
  • Wallis D, Hamblen M, Zhou Y, Venken KJ, Schumacher A, Grimes HL, Zoghbi HY, Orkin SH, Bellen HJ 2003. The zinc finger transcription cistron Gfi1, implicated in lymphomagenesis, is required for inner ear pilus cell differentiation and survival. Development 130: 221–232 [PubMed] [Google Scholar]
  • Wellnitz SA, Lesniak DR, Gerling GJ, Lumpkin EA 2010. The regularity of sustained firing reveals two populations of slowly adapting bear on receptors in mouse hairy skin. J Neurophysiol 103: 3378–3388 [PMC free article] [PubMed] [Google Scholar]
  • Wilson SR, The L, Batia LM, Beattie K, Katibah GE, McClain SP, Pellegrino M, Estandian DM, Bautista DM 2013. The epithelial jail cell-derived atopic dermatitis cytokine TSLP activates neurons to induce crawling. Cell 155: 285–295 [PMC gratuitous article] [PubMed] [Google Scholar]
  • Woo SH, Stumpfova Chiliad, Jensen UB, Lumpkin EA, Owens DM 2010. Identification of epidermal progenitors for the Merkel jail cell lineage. Development 137: 3965–3971 [PMC free article] [PubMed] [Google Scholar]
  • Woo SH, Baba Y, Franco AM, Lumpkin EA, Owens DM 2012. Excitatory glutamate is essential for development and maintenance of the piloneural mechanoreceptor. Evolution 139: 740–748 [PMC free article] [PubMed] [Google Scholar]
  • Woodbury CJ, Koerber Hr 2007. Central and peripheral anatomy of slowly adapting type I low-threshold mechanoreceptors innervating trunk pare of neonatal mice. J Comp Neurol 505: 547–561 [PubMed] [Google Scholar]
  • Wu H, Williams J, Nathans J 2012. Morphologic diversity of cutaneous sensory afferents revealed by genetically directed sparse labeling. eLife 1: e00181. [PMC costless article] [PubMed] [Google Scholar]
  • Xiang Chiliad, Gan L, Li D, Chen ZY, Zhou L, O'Malley BW Jr, Klein Westward, Nathans J 1997. Essential part of POU-domain factor Brn-3c in auditory and vestibular hair cell development. Proc Natl Acad Sci 94: 9445–9450 [PMC gratuitous article] [PubMed] [Google Scholar]
  • Yamashita Y, Ogawa H 1991. Slowly adapting cutaneous mechanoreceptor afferent units associated with Merkel cells in frogs and effects of direct currents. Somatosens Mot Res 8: 87–95 [PubMed] [Google Scholar]

Articles from Cold Spring Harbor Perspectives in Medicine are provided hither courtesy of Cold Spring Harbor Laboratory Press


Tactile Receptors In The Skin,

Source: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4031957/

Posted by: coffeysamot1998.blogspot.com

0 Response to "Tactile Receptors In The Skin"

Post a Comment

Iklan Atas Artikel

Iklan Tengah Artikel 1

Iklan Tengah Artikel 2

Iklan Bawah Artikel