Sunday 18 August 2013

Explanation of localised provoked vulvodynia


Early outlined interdisciplinary theoretical explanation of localised provoked vulvodynia: cortisol/ glucocorticoid receptor distortion and demyelination are the missing links

By:  Journalist Klaus Cort, klauscort@gmail.com

Abstract
Localised provoked vulvodynia (LPV) is unexplained chronic pain in the vulvar vestibule. 10- 16 percent of U.S. women have experienced vulvodynia. Mechanical allodynia and increased intraepithelial innervation in the posterior part of the vulvar vestibule are the signs of LPV. Among vulvodynia-researchers, there is consensus on a multi-factorial aetiology of LPV. Factors statistically associated with and therefore suspected causes of LPV are genes affecting interleukin-1 and mannose-binding lectin, stress, anxiety, depression, use of OCs, repeated vulvovaginal infections, and thereby repeated use of antifungals and antibiotics. However, if these multiple factors produce the same signs, they must affect the same biological mechanisms.
Method and main findings: the approach has been an inter- and multidisciplinary iterative search for and combining of research results, with the aim to explain LPV. Two questions have guided the search and selection:
1) Which biological mechanisms are affected by all the statistically suspected causes of LPV?
2) What can cause the signs of LPV?
The answers found are:
1) The glucocorticoid receptor (GR)-cortisol.
2) Demyelination (accompanied by mast cell degranulation and other inflammatory signs).
Finally, medical literature states that GR-cortisol has a key role in SC myelination/demyelination, mast cell degranulation and in the inflammatory immune response. Thus, GR-cortisol distortion and demyelination links from the statistically suspected causes to the signs of LPV.
Conclusion: LPV is caused by distortion of cortisol metabolism primarily in SCs, which causes demyelination and increased immune response. Demyelination cause innervation of the epithelium and, via neurotransmitters and ionchannels, chronic pain. Interactions between the HPA-axis and the nerve- and immune systems make LPV a systemic condition also affecting the brain.
Consequence: LPV was unexplained. Hormone disturbance by medicine should be minimised.

Introduction
By The International Society for the Study of Vulvovaginal Disease (ISSVD) vulvodynia is defined as “vulvar discomfort, most often described as burning pain, occurring in the absence of relevant visible findings or a specific, clinically identifiable, neurologic disorder.” Vulvodynia is subdivided according to whether the pain is generalised or localised and whether the pain is provoked, unprovoked or mixed (1). Localised provoked vulvodynia (earlier called vestibulodynia or vulvar vestibulitis) and generalised unprovoked vulvodynia (earlier called dysaesthetic vulvodynia, essential vulvodynia or vulvar dysesthesia) seems to be two major subforms. Another subdivision is in primary and secondary, meaning onset of vulvodynia before (or at) and after sexual debut respectively.
Most LPV patients are aged 18 – 35 years, but onset or discovery of primary localised provoked vulvodynia can occur already with the first use of tampon. Two cases where the patient was only 4 years old have been described (2)(3). Generalised unprovoked vulvodynia is most common in menopause or later (4). Mixed forms of vulvodynia do often occur (5). A continuum of signs and symptoms developing or appearing different with age could be an alternative understanding of vulvodynia (6) (and partly (7)). - Advancing age is associated with a higher level of diurnal cortisol secretion, an increased cortisol response to challenge and deterioration of myelin sheaths (8)(9).
The lifetime prevalence of vulvodynia in the USA is 10- 16 percent, 4-7 percent of women have the symptoms at any one time (4)(10). Prevalence is stable among sexually active women of any age. Vulvodynia is rarely diagnosed, but can resolve after a mean duration of 12.5 years (11).
LPV is statistically associated with several other chronic medical conditions: irritable bowel syndrome, chronic fatigue, fibromyalgia, interstitial cystis, stress, anxiety and depression, but the strongest associations are with yeast infections, urinary tract infections and bacterial vaginosis (4)(10)(12)(13)(14)(15). Most results point to a connection between vulvodynia and use of OCs (17)(18)(19)(20)(21)(22).
Increased innervation of the epithelium in the vulvar vestibule is found in nearly all LPV-patients (23)(24)(25). It has been suggested as a diagnostic criterion for localised provoked vulvodynia (25)(26). The vulvar vestibule of LPV patients has no active inflammation (27)(28), but there are presence of inflammatory markers/ immune activity: increased number of mast cells, macrophages and inflammatory cytokines (26)(28)(29)(30)(31).
Women with LPV have an increased superficial blood flow and often an erythema in the posterior parts of the vestibular mucosa, and enhanced systemic pain perception (32)(33). Central sensitization is pronounced in LPV patients with a long history of LPV (34)

Method
The approach has been an inter- and multidisciplinary focus on the problem of the aetiology of LPV, an iterative search for and combining of research results from inside and outside the specific field of vulvodynia research, with the aim to explain LPV. Two simple, and obvious to ask, work-questions has guided the search for and selection of research results:
1) What can cause the signs (mechanical allodynia and neural hyperplasia) of LPV?
2) Which biological mechanisms are affected by all the statistically suspected causes of LPV?
The answers found are mapped and literally outlined in figure 1.

Figure 1 and the text below
When discussing the aetiology of vulvodynia, vulvodynia-researchers most often use the term multi-factorial (31)(35)(36). This is not disagreed upon here. However, if multiple factors produce the same disease, they must act at the same specific biological mechanism in the human body.
Figure 1 depicts many of the factors of LPV, and how they can be related through known biological mechanisms. An arrow, -->,  means “affects” or “cause an effect on”, i.e. OCs affect CBG-level in the blood (in the upper right corner of the figure). Paragraph-headings with arrows refer to the same relations in figure1. Figure 1 is thus a detailed “table of contents”, and depicts the interrelations of the paragraphs below. Paragraphs with no arrow in the heading are background information, selected for better understanding of the explanation presented here and possible relevance for research in and explanation of LPV that goes beyond. A covering review of all the relations depicted as arrows is not the goal with this text, and it should not be expected, as it would require several hundred pages.
 The general table of contents is:
Part I. explains the signs of LPV, chronic pain and neural hyperplasia, as a result of demyelination (arrows in the bottom-right and bottom-mid sections of figure 1).
Part II. Explains demyelination as a result of GR-cortisol distortion (arrow pathways between GR-cortisol and demyelination inside the Schwann cell (SC) in figure 1).
Part III. Explains GR-cortisol distortion as a result of factors statistically associated with LPV (arrow pathways from “Primary causes of LPV” to “GR-cortisol” in figure1).
Part IV. LPV, cortisol and the brain (arrows in the lower left of figure 1, light-blue background)
Part V. Conclusion, significance and consequences.
Please study figure 1 at this point and keep an eye at it while reading the text.





Abbreviations
11β-HSD1: 11β-hydroxysteroid dehydrogenase type 1
11β-HSD2: 11β-hydroxysteroid dehydrogenase type 2
ACTH Adrenocorticotrophic hormone
cAMP. Cyclic adenosine monophosphate
CAR: Constitutive androstane receptor
CBG: Corticoid binding globulin (also called transcortin) 
CORT: Cortisol in humans/corticosterone in mice and rats.
CREB: cAMP response element-binding protein
DHEA: Dehydroepiandrosterone (DHEAS: DHEA-sulphate)
EAAT: Excitatory amino acid transporters
ENaC: Epithelial sodium channel
ERK: Extracellular signal-regulated kinase (ERK1, ERK2)
ER: Estradiol/estrogen receptor
GABA: γ-Aminobutyric acid
GC: Glucocorticoid
GILZ: Glucocorticoid-induced leucine zipper
GITR: Glucocorticoid-induced TNF-receptor-related protein
GR: Glucocorticoid receptor (GRα, GRβ)
GRE’s Glucocorticoid response elements
hGR: Human Glucocorticoid receptor (hGRα, hGRβ)
IL: Interleukin (IL-1α, IL-1β, IL-6)
IL-1RA: Interleukin-1 receptor antagonist
JNK: c-Jun N-terminal kinase
LPV: Localised provoked vulvodynia
LRP-1: Low Density Lipoprotein Receptor-Related Protein 1
MAPKs: Mitogen-activated protein kinases (ERK, JNK, p38 MAPK etc.)
MBL: Mannose binding lectin
MR: Mineralocorticoid receptor
NCX: Na(+)/Ca(2+) exchanger, sodium-calcium exchanger
NF-κB: Nuclear factor-kappaB
NGF: Nerve growth factor
NMDA: N-methyl-D-aspartate
NMDAR: N-methyl-D-aspartate receptor
P0: Protein zero
PI3K: Phosphatidylinositol 3-kinase
PKC: Protein kinase C (PKC-α, -β , -δ and –ζ)  
PMP22: Peripheral myelin protein-22
PNS: Peripheral nervous system
PXR: Pregnane X receptor
RAR: Retinoic acid receptor
RVVIs: Recurrent vulvovaginal infections
RXR:  Retinoid X receptor
SGK: Serum- and glucocorticoid-inducible kinase (SGK1, SGK2, SGK3)
SC: Schwann cell
SHBG: Sex hormone binding globulin
TNF: Tumor necrosis factor (TNF-α)
TRP: Transient receptor potential (TRPV1, TRPA1, TRPM8)
TRPA1: Transient receptor potential ankyrin 1
TRPV1: Transient receptor potential vanilloid 1



I. Demyelination causes the signs and symptoms of LPV

This part describes how demyelination changes sensory-nerves into unmyelinated (of course) and pain transmitting nerves with a propensity to grow - in the case of LPV into the epithelium of the vulvar vestibule.

Demyelination ­­-->  neural hyperplasia
Myelin contains several proteins, which inhibit or restrict neural growth (i.e. myelin-associated glycoprotein (MAG) and neurite outgrowth inhibitor (Nogo)), and demyelination cause neuronal growth (37)(38)(39). Loss of myelin may thus explain the innervation of the epithelium observed in LPV.
Neural hyperplasia associated with pain and an increased number of mast cells is also found in non-inflamed appendices from patients with acute appendicular pain (40)(41). (Several factors causing both demyelination and neural hyperplasia are described in part II, and demyelination in the brain of LPV patients is shortly described in part IV.).

Demyelination --> Ion-channels
Demyelination causes altered axon-SC interactions: axonal components of nodes fragment and disappear, glial and axonal paranodal and juxtaparanodal proteins no longer cluster, and axonal Kv1.1/Kv1.2 K+ channels move from the juxtaparanodal region to appose the remaining heminodes (42). Demyelination trigger membrane remodelling in injured afferents and perhaps in uninjured neighbours, which causes increased cellular excitability: enhanced membrane resonance, rhythmogenesis, and ectopic spiking, which are the characteristics of a primary neuropathic pain signal. This is due in large part to subtype-selective abnormalities in the expression and trafficking of Na+ channels (43). Na+ channel isoforms are differentially targeted to distinct domains of the same axon in a process associated with formation of compact myelin. During development, Na(v)1.2 is expressed first and becomes clustered at immature nodes of Ranvier, but as myelination proceeds, Na(v)1.6 replaces Na(v)1.2 at nodes (44). Demyelination causes a significant switch from Nav1.6 to Nav1.2 expression (45). SC remyelination restores the normal pattern of Nav1.6 and Kv1.2 at nodes of Ranvier (46).
These results might in addition suggest that myelin has a key role in keeping homeostatic concentrations of Na(+) and K(+) at nodes of Ranvier.

Demyelination --> TRPV1, NMDA --> chronic pain
Focal peripheral nerve axon demyelination is accompanied by nociceptive pain behaviour in mice. The demyelination leads to delayed functional expression of neuronal chemokine receptors. Chemokine signalling by both injured and adjacent, uninjured sensory neurons are correlated with the maintenance phase of a persistent pain state. Chemokines can directly excite subsets of sensory neurons. This excitation is likely to be due to transactivation of ion channels, such as the transient receptor potential vanilloid 1 and transient receptor potential ankyrin 1 (TRPV1 and TRPA1), expressed by sensory nerves (47). Cortisol has been found to regulate some chemokine receptors (48)(49). TRPV1 and TRPA1 channels are members of the TRP superfamily of structurally related, non-selective cation channels. The functions of TRPV1 and TRPA1 interlink with each other to a considerable extent, especially in relation to pain and neurogenic inflammation where TRPV1 is co expressed on the vast majority of TRPA1-expressing sensory nerves (50). Increased TRPV1 innervation in vulvodynia tissues compared
TRPV1 activation induces Ca(2+) entry, a prolonged elevation of presynaptic mitochondrial and cytosolic Ca(2+) and a concomitant enhancement of glutamate release at sensory synapses and action potential firing by postsynaptic neurons (52).
Activation of N-methyl-D-aspartate (NMDA) receptors (NMDAR) sensitizes TRPV1 by enhancing serine phosphorylation through protein kinase C (PKC). Thus it seems that the NMDAR and TRPV1 forms a signalling complex that underlies the sensitization of nociceptors (53)
Models of neuropathic pain are created by inflicting injuries to peripheral sensory nerves i.e.  chronic constriction, axotomy and demyelination. Demyelination can be caused by increased Ca(2+) in SCs and is accompanied by increased neuronal Ca(2+) (47)(54).
Axotomy, on the other hand, causes loss of neuronal inward Ca(2+) flux through voltage-gated Ca(2+) channels and decreased neuronal cytosolic Ca(2+) (55).
Apart from urging caution in the interpretation of these models, this might also be relevant in the understanding of the effectiveness of vestibulectomy as a treatment of LPV. Testosterone and progesterone are other treatments of LPV, and dehydroepiandrosterone (DHEA) might have a potential for the same purpose, as all three hormones block Ca(2+) channels of varying types (22)(56)(57)(58)(59).
Capsaicin, the pungent ingredient in hot chilli peppers activates TRPV1, which leads to a burning sensation (60).
In the USA, Hispanic women are 80% more likely to experience chronic vulvar pain than are White and African American women (61).
Is it because of differences in chilli intake, differences in genes or more stress in Hispanic women (maybe caused by minority/immigrant-situation and/or Hispanic sex-roles)?
Whereas low capsaicin concentrations results in sensitization and activation of TRPV1 receptors, higher concentrations of topical capsaicin can result in desensitization of TRPV1-positive afferents and eventually withdrawal of epidermal nerve fibres (62)(63). Capsaicin could thus be both a cause and a treatment possibility in LPV.

Ion-channels --> chronic pain
TRPV1 is of course not the only ion-channel involved in pain sensation. Other TRP-channels, Voltage-gated Ca(2+) channels and Voltage-gated K(+) channels also play major roles in the development and maintenance of neuropathic pain  (64)(65)(66)(67). The importance of the Na(+)/Ca(2+) exchanger (NCX) for both neuronal and glial cells is described in part II., because NCX’s are regulated by some of the same factors as, and interacts with, myelination.

GR-cortisol --> ion-channels
Corticosterone, injected or induced by water avoidance stress, leads to increased TRPV1 receptor expression and function in rats (68).
GR-cortisol regulates the cross membrane exchangers of Na(+) for Ca(2+), K(+) and H(+) respectively(69)(70)(71)(72). In addition, the Na(+) transport by the epithelial sodium channel (ENaC) is regulated by glucocorticoids (GCs) via  GC-induced leucine zipper (GILZ) and serum- and GC-inducible protein kinase 1 (SGK1). An ENaC-like channel has recently been found in rat PC12 cells (a neuronal cell model) and in a human colonic cell line. Thus, ENaC channels are most likely to be present in mucosal tissue and in the nerves herein. GILZ expression is also rapidly stimulated by aldosterone, which strongly stimulates ENaC-mediated Na(+) transport by inhibiting extracellular signal-regulated kinase (ERK) signalling. However, the GR is indispensable for physiological responses to aldosterone in ENaC induction via the mineralocorticoid receptor (MR), and SGK1 and ERK interact. (73)(74)(75)(76)(77).
GC-induced hypertension is in part caused by a dysregulation of Na(+) homeostasis (78).
Women with LPV have an increased superficial blood flow in the posterior parts of the vestibular mucosa most probably caused by a neurogenic vasodilatation (32) - being a result of oppositely  dysregulated Na(+) homeostasis (hypotension), it is implied here.
In a randomized, double blinded, placebo-controlled study (of  botulinum toxin A) 0.5 mL saline, with and without botulinum toxin A, injected in the musculus bulbospongiosus produced equal and significant pain reductions (P < 0.001) in LPV patients (79).
Increased activity or even re-reversing of the Na(+)/Ca(2+) exchanger (NCX) could be the underlying effect of this unintended treatment.
If there is strong (local?) lack of Na(+) even the placebo-concentration has an alleviating effect.
Na(+) regulates different glutamate receptors outside and inside the neuronal cell membrane. The enhancement of NMDARs by intracellular Na(+) interacts with Ca2(+)-dependent inactivation (80).     NaCl (or another Na(+) source) as a conservative treatment of LPV should be considered further. A less conservative treatment of LPV could be a mineralocorticoid aimed at increasing GILZ.

GR-Cortisol  --> Neurotransmitters
Glutamate is one of the major excitatory neurotransmitters in the central nervous system, but has also a role in the transduction of sensory input in the peripheral nervous system (PNS), and in particular in the nociceptive pathway. There is strong support for the presence of GRs on presynaptic nerve terminals acting to facilitate the release of neuronal glutamate (81).
Serotonin is regulated by MR and GR responsive promoter elements (82)(83)(84).
A PNS interconnection between GR-cortisol and dopamine, which involves both SCs and neurons, is described  in “GR-cortisol à arylsulphatase A à demyelination” in part II.

II. GR-cortisol distortions cause demyelination

All the molecular instruments playing the symphony of myelination/demyelination are conducted by GR-cortisol. If the conductor is disturbed, you are sure of a bad concert. However, if some of the instruments play out of tune, it can also result in a bad performance – demyelination.
Of course, the way the molecular instruments play affects the GR-conductor and the other molecular instruments. Only for limitation-reasons, these relations are not, or only briefly described here.

The GR and myelin genes
The regenerative (demyelinating) response of SCs is directly related to the pathophysiology of a number of neurodegenerative diseases, and is dependent on an intricate gene regulatory program coordinated by a number of transcription factors and microRNAs, which are correlated with myelination and proliferation gene clusters (85)
Of key importance is the mutually antagonistic relationship between Early growth response protein 2 (Egr2 / Krox20) and the transcription factor c-jun that regulates the transitions between nonmyelinating and myelinating SCs. This antagonistic relationship is regulated by GILZ (86)(87).
The promotors of peripheral myelin protein-22 (PMP22) and protein zero (P0) genes are only activated in SCs, and only by ligand activated GR. Strangely, the GR antagonist RU486 does not abolish the effect of glucocorticosteroids, instead it stimulates promoter activities by itself (88). In SCs, the GR also makes use of unusual coactivators for its binding to the GC response elements (GRE’s). Expected coactivators inhibits GR transcriptional activity, which in stead is mediated by β-catenin (89)(90).

Neuregulin-1
Neuregulin-1 (Nrg1) provides a key axonal signal that regulates SC proliferation, migration and myelination through binding to SC receptors (called ErbB2/3). Both the membrane-bound type III and the soluble isoform II of Nrg1 elicit a promyelinating effect at low concentrations, and they both inhibit myelination at higher concentrations, by activation of mitogen-activated protein kinases (MAPKs) and induction of increased expression of the transcription factor c-Jun (91)(92). However, Nrg1 type I expression in SCs themselves plays a pivotal role in remyelination (93).

GR-cortisol --> demyelination
Cortisol and progesterone are decisive in the production of myelin. The two steroids both initiate and control the rate of myelin formation (94). A single SC produces myelin equivalent to many thousands of its own weight. The number of GC receptors (GR’s) is therefore extremely high in schwann-cells, and schwann-cells are extremely sensitive to variations in cortisol level in a u-shaped manner. A small increase relative to the normal physiological level can be beneficial, while both decreases and lager increases in cortisol level can cause endoplasmic reticulum stress, wrongly folded proteins, myelin-failure and eventually schwann-cell death (95)(96)(97).





GR-cortisol --> PXR/… --> demyelination
When bound to GR, cortisol regulates the activity of retinoic acid (RA) bound to its receptors RAR and RXR (98). RA-RAR regulates the production of proteins essential for myelin. RA up-regulates myelin basic protein (Mbp) and myelin P0, when connecting to RXR, and it down-regulates the production of MAG when connecting to RAR. Changes in GR-cortisol level can therefore lead to myelin failure (99). RA, acting through the RAR-β, inhibits the neuronal membrane-bound receptor of myelin-activated Nogo, through the transcriptional repression of Nogo receptor interacting protein (Lingo-1), which results in lacking inhibition of neurite outgrowth (100).
Lingo-1 is a potent inhibitor of oligodendrocyte differentiation and myelination, both when expressed by oligodendrocytes and when expressed by neuronal cells (101).
At least 70 percent of myelin is lipids. The lipid metabolism is vulnerable in part due to the particular lipid composition of myelin and the transport of lipid-associated myelin proteins (102).
The production of lipids is regulated by the pre-hormone pregnenolone, when this connects to its receptor PXR (pregnane X receptor). PXR-pregnenolone is regulated by GR-cortisol. Changes in GR-cortisol level can therefore lead to failures in lipid metabolism (98).
A GC-controlled gene network is involved in the regulation of triglyceride homeostasis (103). This was found in adipocytes, but might apply for other cells with high production/maintenance of triglycerids like SCs.

GR-cortisol --> kinases --> neural hyperplasia, demyelination --> pain
Kinases are enzymes that can rapidly and reversibly phosphorylate specific residues of cellular proteins and as such affect their structure, function, location or metabolism (78).
It is well documented that MAPK pathways can increase peripheral pain sensitivity (104).
Activation of both ERK and p38 MAPK signalling pathways are involved in neurite outgrowth and differentiation of PC12 cells toward a neuronal phenotype (105). Normally, after nerve injury, SCs dedifferentiate into a progenitor-like state, proliferate, and repopulate the damaged nerve. Once axons have regenerated SCs then redifferentiate and remyelinate. Elevated MAPK (/ERK) signalling in SCs is a crucial trigger for SC dedifferentiation in vivo (106)(107). Both inhibition and activation of p38 MAPK cause demyelination (108)(109)(110)(111), which suggests a U-shaped relation, when demyelination is depicted as a function of p38 MAPK activity. GR-Cortisol activates MAPK (/ERK) through genomic mechanisms, but also interacts with MAPKs in a non-genomic way (112).
PKC-mediated phosphorylation of  the myelin protein P0 is necessary for P0-mediated homophilic adhesion, and alteration of this process can cause demyelinating neuropathy in humans (113).
PKC phosphorylates the transcription factor Sp1 that can activate the myelin basic protein (MBP) promoter (114)
PKC is a key component in the signalling pathways that mediate the inhibitory activities of myelin on neuronal growth. MAG, Nogo and oligodendrocyte myelin glycoprotein (OMgp) all interact with the same receptor complex to effect inhibition via PKC (115)(116).
GR activation increase mRNA and protein level of PKC. However, this effect is isoform specific. The PKC isoforms -α, -β and -ε are strongly increased, while the -δ and -ζ  isoforms are not affected.  In mesenteric arteries from hypertensive rats Dexamethasone decrease PKC activation (117)(118(119), which suggests an inverted U relationship between GR activation (x-axis) and PKC activity (in the hypertensive rats GR-activity and PKC-activity is all ready near or beyond the turning point in the inverted U before the dexamethasone treatment, as hypertension is regulated by GR).
NMDA receptors and PKCγ are regulated by GR through a cyclic adenosine monophosphate (cAMP) response element-binding protein (CREB)-dependent pathway (shown in spinal cord by chronic morphine exposure)(120).

Glutamate, Ca(2+) --> demyelination, pain
Activation of myelinic NMDA glutamate receptors mediates Ca2+ accumulation in central myelin in response to chemical ischemia in vitro. Given that axons are known to release glutamate, this suggests a mechanism of axo-myelinic signalling of importance for disorders in which demyelination is a prominent feature (121).
Neurotransmitters (i.e. γ-Aminobutyric acid (GABA), acetylcholine, adenosine, glutamate) are active on SCs. Prevention of glutamate induced excitatory, toxic and demyelinating effects is desirable to preserve the integrity of PNS. Removal of glutamate from the extracellular space is accomplished by the high affinity glutamate transporters called excitatory amino acid transporters (EAATs), which are present in SCs, in the myelin layer and at neuronal synapses (122)(123(124).
The glutamate released by neurons into the synaptic cleft is inactivated by EAATs that catalyzes the co-transport of 3 Na(+) ions, one H(+) ion, and one glutamate molecule into the cell, in exchange for one K(+) ion. Five EAATs has been identified:EAAT1 (also called GLAST), EAAT2 (GLT-1), EAAT3 (EAAC1 or excitatory amino acid carrier 1), EAAT4 and EAAT5. EAATs are affected by several effectors, including free radicals, arachidonic acid, protein kinases A, B and C, phosphatidylinositol 3-kinase (PI3K)  and SGK1, SGK2, and SGK3. All three SGK isoforms and PKB increase EAAT2 activity and plasma membrane expression. PKC activation on the other hand leads to the internalization of both EAAT2 and the dopamine transporter, and thereby a reduction in neurotransmitter clearance capacity (125)(126)(127)(128)(129).

NCX and EAAT
The sodium-calcium exchanger NCX, of which there three isoforms are known, is a bidirectional transporter that catalyzes the electrogenic exchange of 3 Na(+) for 1 Ca(2+), depending on the electrochemical gradient of the substrate ions (130). During oligodendrocyte precursor cells (OPC) differentiation into oligodendrocyte phenotype NCX1 is downregulated and  NCX3 is strongly upregulated. NCX3-knockout mice show reduced size of spinal cord and marked hypo-myelination, as revealed by decrease in myelin basic protein (MBP) expression and increase in OPC number (131)
NCX1 and NCX3 basal expression decreases when c-Jun N-terminal kinase (JNK) or ERK 1/2 are blocked. Whole-cell Na(+) /Ca(2+) exchange decreases when JNK and ERK 1/2 are blocked and increases when MAPKs are activated by nerve growth factor (NGF) (132).
NCX1 and NCX3 up-regulation contribute to the survival action of the PI3K pathway (during chemical hypoxia) (133).
Both PKC and PKA activation enhance NCX reverse mode, which in neurons results in Ca(2+) influx and NMDA excitotoxicity (134)(135).
EAAC1 (also called EAAT3), is expressed in neuronal and glial mitochondria where it participates in glutamate-stimulated ATP production. There is  colocalization, mutual activity dependency, physical interaction between EAAC1 and NCX1 both in neuronal and glial mitochondria, and NCX1 is the key player in the glutamate-induced energy production (136).
During ischemia the mitochondrial Na(+)/Ca(2)+ exchanger, driven in the Na(+) import/Ca(2+) export mode, contributes to Ca(2+) increase in the cytosol (in rat optic nerve) (137).
Thus, assuming the presence of the NCX-EAAT-complex in the cell membrane, the normal NCX mode (Na(+) import, Ca(2+) export) competes with EAAT for the 3 Na(+), which regulates glutamate homeostasis. Excessive glutamate release and /or lack of extracellular Na(+) makes EAAT win this competition, NCX is forced into reverse mode, the NA (+) is recycled, and the cost of removing excessive glutamate is increased intracellular Ca (2+). In SCs and myelin, the increased intracellular Ca(2+) leads to demyelination. In neuronal cells, it leads to excitotoxicity.
However, NCX is not the only source of intracellular Ca(2+). TRPV1 has been mentioned earlier.
Another example: Increased intracellular calcium causes functional derangement in SCs from rats with Charcot-Marie-Tooth neuropathy (PMP22 gene overexpression). A PMP22-related overexpression of the P2X7 purinoceptor/channel (members of the family of ionotropic ATP-gated receptors) leads to the influx of extracellular Ca(2+) and demyelination (138).

GR-cortisol --> LRP-1 --> NMDA
Low Density Lipoprotein Receptor-Related Protein 1 (LRP-1) modulates NMDA receptor-dependent intracellular signalling and NMDA-induced regulation of postsynaptic protein complexes (139).
Deletion of  the LRP1 gene in SCs (scLRP1(-/-)) induces abnormalities in axon myelination and in ensheathment of axons by nonmyelinating SCs in Remak bundles. These anatomical changes in the PNS are associated with mechanical allodynia, even in the absence of nerve injury, and central sensitization in pain processing including increased p38MAPK activation and activation of microglia in the spinal cord (140).
LRP1 is regulated by ligand activated GR (141)(142).
LRP1 functions as a potent activator of PI3K in SCs and, by this mechanism, increase the SC unfolded protein response, which limits apoptosis (143).

GR-cortisol --> arylsulphatase A, dopamine --> demyelination
In humans, most Dopamine circulates as dopamine sulfate, which can be de-conjugated to bioactive dopamine by arylsulfatase A. Human adipocytes express functional dopamine-receptors and arylsulfatase A, suggesting a regulatory role for peripheral dopamine (144).
Dopamine receptors D1 and D5 (D1-like receptors), are linked to a stimulatory G protein, that stimulate adenylyl cyclase and increases cAMP production. D2R, D3R, and D4R (D2-like receptors) are linked to an inhibitory G protein, that inhibit adenylyl cyclase and calcium channels, and modulate potassium channels (145).
Metachromatic leukodystrophy is a lysosomal storage disorder caused by deficiency in the sulfolipid degrading enzyme arylsulfatase A. In the absence of a functional arylsulfatase A, gene sulpholipids accumulate. The storage is associated with progressive demyelination and various finally lethal neurological symptoms. Lipid storage, however, is not restricted to myelin-producing cells but also occurs in neurons. Accumulation in neurons contributes to disease phenotype, hyperexcitability and axonal degeneration (146).
Cortisol in physiological concentrations (0.03 microM) causes an increased accumulation of myelination-associated sulpholipids in SCs. It is caused by a cortisol-concentration-dependent inhibition in arylsulphatase A activity (147)(148).
High cortisol levels may thus cause a “metachromatic leukodystrophy-light”: dys-/demyelination, low arylsulphatase A  leading to missing stimulation of peripheral neuronal dopamine receptors and hyperexcitability.

GR-cortisol --> microRNA --> neural hyperplasia, demyelination
MicroRNAs are small non-coding RNA molecules, which functions in transcriptional and post-transcriptional regulation of gene expression. A cohort of microRNAs coordinate SC dedifferentiation through a combinatorial modulation of their positive and negative gene regulators during the acute phase of PNS injury (149)(150).
MicroRNA-221 and -222 promote SC proliferation and migration after sciatic nerve injury (151). MicroRNA-222 promotes neurite outgrowth from adult dorsal root ganglion neurons following sciatic nerve transaction (152).
RNA polymerase II  is an enzyme that catalyzes the transcription of DNA to synthesize precursors of mRNA and most microRNA (153)(154) The GR functions at multiple steps during transcription initiation by RNA polymerase II (155).

Other steroids --> demyelination
Progesterone and its derivatives also control the production of proteins that are unique and essential for myelin. The gene expression of Glycoprotein Po is stimulated via the progesterone receptor. In addition, tetrahydroprogesterone increases PMP22 gene expression via the GABA-A receptor. Also over expression of PMP22 can cause dysmyelination. Both lack and surplus of progesterone can therefore cause dys- and demyelination. SCs can produce progesterone (156)(157)(158).
Testosterone is vital for the production of P0 and PMP22. When connecting to the androgen receptor, testosterone controls P0, while the control over PMP22 is most likely via the GABA-A receptor (159).

Other steroids --> GR-cortisol
Key regulators of cortisol activity are the enzymes 11β-hydroxysteroid dehydrogenases, 11β-HSD1 and 11β-HSD2. 11β-HSD1 reduces cortisone to active cortisol. 11β-HSD2 oxidizes cortisol to the inactive cortisone. 11β-HSD2 is strongly expressed and active in quiescent (myelinating) SCs. In proliferating SC, 11β-HSD2 exhibits a strong decrease in activity and mRNA concentration. Metabolites of progesterone affects cortisol metabolism by inhibiting 11β-HSD1 and 11β-HSD2. A metabolite of DHEA, competitively inhibits 11β-HSD1(160)(161)(162)(163).


III. The statistical suspects of LPV cause GR-cortisol distortion

The GR has been shown by microarray analysis to regulate up to 10–20% of the human genome in different cell types (164). The effect of human GR (hGR) antagonists is worrying and is likely to result in adverse effects (98). Moreover, logically the same goes for agonists of hGR. The GC function is the most important regulatory system of homeostasis (165).

Oral contraceptives --> CBG, SHBG --> GR-Cortisol
Endogenous pain modulation of experimentally induced (acute) pain is less effective in users of oral contraceptives (OCs) than in normally menstruating women (166).
OCs induce significant lower mechanical pain thresholds in the vestibular mucosa in healthy women. The most sensitive area is the posterior vestibule, the by far most common localisation of LPV. OCs might thus be one contributing factor in the development of LPV (20).
OCs increases plasma concentrations of corticoid binding globulin (CBG also called transcortin) and  sex hormone binding globulin (SHBG) in a magnitude of 50-300 percent. The CBG increase results in increased total cortisol, but unchanged free serum cortisol. However, women on oestrogens may have altered free serum cortisol kinetics and they thus may be potentially overexposed to GCs. The SHBG increase results in decrease of free testosterone and other androgens (167)(168)(169)(170)(171)(172).
CBG may regulate access of GCs to the brain and other tissues of the body. CBG is expressed in the human hypothalamus and cerebrospinal fluid. CBG functions as a protein thermocouple that is exquisitely sensitive to temperature change, releasing cortisol in response to increasing temperatures within the human physiological range (173).

Genes and CBG
Some CBG-null mice have an 10-fold increase in free corticosterone levels other have markedly reduced total circulating corticosterone at rest and in response to stress (174)(175).
In humans, a mutation in the CGB-gene is associated with hypotension and fatigue. The CBG null patients have normal free serum cortisol levels but lack a CBG-bound pool of readily releasable cortisol (168)(176). Two other genetic variants of CBG, the Leuven and Lyon mutations, reduce CBG cortisol binding affinity 3- and 4-fold, respectively (177). Chronic fatigue is co-morbid with LPV. This raises the question: are mutations in the CBG-gene a cause of LPV?

Oral contraceptives --> SHBG --> LPV
SHBG may also play a role in generating the LPV-like reduced pain thresholds found in the posterior vulvar vestibule of (otherwise) healthy OC-users. Among LPV patients using different combined contraception, those using low dose estradiol and second generation progestin have significantly lower increase in SHBG levels, that is associated with less reduced free total testosterone ratios and less sexual pain (178).

Oral contraceptives --> other steroids --> GR-cortisol
However, the main effect of OCs is to increase free estradiol and progestin (although  some will bind to the increased SHBG and CBG, respectively).
The ER-estradiol interacts with the DNA-binding transcription factor c-Jun, which promotes the nonmyelinating state of SCs. Estrogen can trigger rapid ‘‘non-genomic signalling’’ associated with the activation of second messengers as the MAPK, PKC and PI3K which can cause increased intracellular Ca(2+) and demyelination (179).
Estradiol is an antagonist of GC-induced GILZ gene expression in human uterine epithelial cells and murine uterus. GILZ gene expression is associated with several of the immune-related functions of GCs (180) - and myelination and Na(+)-homeostasis, as mentioned earlier.
17-β estradiol produces significant decreases in GR concentrations and GR mRNA levels. Chronic E(2) treatment reduce GR to very low levels. The estrogen mediated suppression is long lasting (more than 10 days after withdrawal) and can not be easily reversed. (in MCF-7 breast cancer cell line) (181).
Estrogen agonists down regulate GR through an ER-dependent increase in Mdm2 protein, an E3 ubiquitin ligase that targets the GR to the proteasome (in MCF-7 breast cancer cell line) (182).
Estrogen reduce ligand-induced GR phosphorylation, which is associated with the active form of GR, by increasing expression of protein phosphatase 5, which mediates the dephosphorylation of GR at Ser-211 (in MCF-7 breast cancer cell line) (183).
Estradiol causes a dysregulation of HPA axis negative feedback as evidenced by the inability of dexamethasone to suppress diurnal and stress-induced CORT (Cortisol in humans/corticosterone in mice and rats) and ACTH secretion. The ability of estradiol to inhibit GC negative feedback occurs specifically via ER-α acting at the level of the paraventricular nucleus of the hypothalamus (184). (The respectively referenced authors’ choice of word for the female sex hormone(s) has been used).

Antifungals ­­--> GR-cortisol --> CAR, PXR
Imidazole antimycotic drugs possess GC antagonist activity by virtue of occupancy of GC receptor sites. Dose-dependent, competitive displacement of [3H]dexamethasone binding is in the potency sequence: clotrimazole > ketoconazole > RS 49910 (185).
Ketoconazole and miconazole are antagonists of hGR and inhibits the expression of GR-responsive genes: tyrosine aminotransferase and both PXR and CAR, and further CAR and PXR target genes including cytochromes P450: CYP2B6, CYP2C9, and CYP3A4. Fluconazole has no such effects (98). CYP3A4 is involved in the hydroxylation and termination in activity of steroid hormones, especially testosterone, estrogen and cortisol (186).
Ketoconazole and erythromycin causes dramatic conformational changes upon binding to CYP3A4, a differential but substantial (>80%) increase in the active site volume, providing a structural basis for ligand promiscuity of CYP3A4 (169). Clotrimazole induces overexpression of PXR (187)(188).

Antifungals ­­--> CAR, PXR --> lipid metabolism
Three triazoles used in agriculture myclobutanil, propiconazole and triadimefon all significantly perturb the fatty acid, steroid, and xenobiotic metabolism pathways in the male rat liver. In addition, triadimefon modulate expression of genes in the liver from the sterol biosynthesis pathway. The three triazoles perturb fatty acid and steroid metabolism predominantly through the CAR and PXR signalling (189).

Antifungals --> Other steroids
Imidazoles (econazole, ketoconazole, miconazole, prochloraz) and triazoles (epoxiconazole, propiconazole, tebuconazole) all show endocrine disrupting effects. The mechanism seems to be disturbance of steroid biosynthesis. The conazoles decrease the formation of estradiol and testosterone, and increase the concentration of progesterone, indicating inhibition of enzymes involved in the conversion of progesterone to testosterone (190).

Clotrimazole --> GR-cortisol --> pain
Clotrimazole is a widely used drug for the topical treatment of vaginal yeast infections. Common side effects of topical clotrimazole application include irritation and burning pain of the skin and mucous membranes. Transient receptor potential (TRP) channels in primary sensory neurons underlie these unwanted effects of clotrimazole. The transient receptor potential (TRP) superfamily is a large group of cation channels that play a general role as cellular sensors of thermal, mechanical and chemical stimuli, and in the initiation of irritation and pain caused by such stimuli. Clotrimazole in clinically relevant concentrations is an agonist of TRPV1 and TRPA1 and a potent antagonist of TRPM8. A covalent binding of clotrimazole to the channels is unlikely (191).
Clotrimazole’s strong competitively displacement of cortisol from the GR and overactivation of both PXR and TRPV1, suggests that clotrimazole overactivates the GR, which is unlike other
–azoles that block the GR and are antagonists of TRPV1 (98)(185)(192)(193).

Undue use of medicine --> antibiotics, antifungals
When women present to physicians, yeast vaginitis is often diagnosed solely based on self-diagnosis or the patient’s history - even though an accurate diagnosis requires clinical assessment, a positive fungal culture result, and a vaginal pH assessment. Among women who use over-the-counter antifungal medications, 50% do not have yeast infections (194).

Chronic pain --> Painkillers --> GR-cortisol
Painkillers are not normally included on the list of “statistical suspects” of LPV. However, chronic pain and Hypersensibility may lead to more than normal consumption of painkillers, especially when LPV is not diagnosed. As painkillers also interfere with cortisol metabolism, they are here added to the list.
Acute adrenocorticotrophic hormone (ACTH)-mediated cortisol production in trout interrenal cells in vitro is significantly depressed (20-40%) by salicylate, ibuprofen and acetaminophen. Salicylate is the major metabolite and active component of aspirin (acetylsalicylic acid). Salicylate is a corticosteroid disruptor in trout and the targets include the key rate-limiting step in interrenal steroidogenesis and brain GC signalling (195). 
Sodium salicylate significantly enhances neuronal excitation in the hippocampal CA1 area of rats. Aspirin might impair hippocampal synaptic and neural network functions through its actions on GABAergic neurotransmission. Given the capability of aspirin to penetrate the blood-brain barrier, this implies that aspirin intake may cause network hyperactivity and be potentially harmful in susceptible subpopulations (196).
The function of Type II GC receptors is inhibited by sodium salicylate in a non- competitive way. Sodium salicylate enhances the density of Type III GC receptors. Depending on the dosage, sodium salicylate increase the number of sites of binding 3H-corticosterone to type III GCal receptors (197)(198).
(In this older nomenclature of GC receptors: type I= MR, type II= GR and Type III= CBG-like, likely to be the 11β-HSD enzymes (199)(200)).

Painkillers --> other steroids
Mild analgesic drugs have been associated with anti-androgenic effects in animal experiments. Intrauterine exposure to mild analgesics is a risk factor for development of male reproductive disorders. There is an association between the timing and the duration of mild analgesic use during pregnancy and the risk of cryptorchidism. These findings are supported by anti-androgenic effects in rat models leading to impaired masculinization, a reduction in the anogenital distance (201). 

Antibiotics --> GR-Cortisol
Penicillins and cephalosporins have a strong action on the most important regulatory system of homeostasis, the GC function. Penicillin G and cefazolin induce a dose-dependent increase in the density of the type III GC receptors and a decrease in the affinity of 3H-corticosterone with the type III GC receptors. The activation of the function of the type III GC receptor by penicillin G and cefazolin is not competitive. Cefazolin also increase the density of the type II GC receptors (165). A combination of trimethoprim and sulfamethoxazole induce a rapid physiological stress response, an increase in plasma cortisol and glucose, and sensitivity, which requires more than 48-h period for regaining homeostasis, in two species of fish (202). Trimethoprim and the combination with sulfamethoxazole are often used in treatment of urinary tract infections, a condition that is co-morbid with LPV (12)(203).

Antibiotics --> GR-cortisol --> NMDA --> pain
LPV and use of antibiotics against vulvovaginal infections are statistically associated (13). Neurotoxicity is common among many groups of antibiotics in at-risk patients and can range from ototoxicity, neuropathy and neuromuscular blockade to confusion, non-specific encephalopathy, seizures and status epilepticus. The underlying mechanism of neurotoxicity is in many patient-cases activation of NMDA receptors and/or inhibition of GABA-A receptors (204).
MAPKs contribute to central sensitization and neuropathic pain. A particular chain of events resulting in NMDA activation (shown in the superficial spinal cord) is:
p38 MAPK --> chemokine CCL2 --> TNF-α -->  NMDA  --> pain.
The proinflammatory cytokines, that at low concentrations (1–10 ng/ml) induce central sensitization, are tumor necrosis factor (TNF)-α , interleukin (IL)-1β and IL-6. TNF-α enhances excitatory synaptic transmission by increasing the frequency of spontaneous excitatory postsynaptic currents and the amplitude of AMPA- or NMDA-induced currents. IL-6 inhibits inhibitory synaptic transmission by reducing the frequency of spontaneous inhibitory postsynaptic currents and the amplitude of GABA- and glycine-induced currents. IL-1β can both enhance excitatory synaptic transmission and reduce inhibitory synaptic transmission (205)(206)(207).
Cortisol activates MAPK through genomic mechanisms, but also interacts with MAPK in a non-genomic way. Cortisol is known to regulate some chemokine receptors. GC-induced TNF- receptor-related protein (GITR) is - not surprisingly - induced by cortisol. Cortisol affects NMDA and GABA signalling (48)(49)(111)(112)(208).

Metronidazole, chromatin, cancer and demyelination
Metronidazole (nitroimidazole) is an antibiotic often used against vulvar infections and Crohn’s disease. Metronidazole is reasonably anticipated to be a human carcinogen based on sufficient evidence of carcinogenicity from studies in ex­perimental animals (209). Although options are limited, alternative therapies to the nitroimidazole antibiotics are available (210).
In women, there is an association between bacterial vaginosis and cervical intraepithelial neoplasia (211). Crohn’s disease is a recognized risk factor for cancer of the small intestine, with relative risks reported as high as 60. A meta-analysis showed a relative risk of 33.2 (95% CI: 15.9-60.9) (212). Patients with Crohn's disease also have an increased risk of colorectal cancer (213).  Whether the increased risk of cancer in the two patient-populations is caused by metronidazole or the two different inflammation-related diseases cannot be decided because of the “experimental setup” in this in vivo big scale test in humans.
Nitroimidazole derivatives exhibit genotoxic effects, large numbers of sister chromatid exchanges and chromosomal aberrations in cultured human lymphocytes and in the non-human primate, Cebus libidinosus. These effects are not randomly distributed, but concentrated at chromosomes rich in heterochromatin (214)(215)(216)(217).
The GR regulate the activity of many genes by binding to target sites within promoter regions of genes assembled as chromatin. Transcriptional activation is mediated by the GR remodelling of chromatin complexes (218)(219)(220). All major human cancers, in addition to having a large number of genetic alterations, exhibit prominent epigenetic abnormalities that can be used as biomarkers for the molecular diagnosis of cancer (221).
Metronidazole is associated with numerous neurologic complications, most commonly peripheral neuropathy. Case reports have been published describing motor, sensory and autonomic neuropathy. Nerve conduction studies demonstrate a peripheral neuropathy manifested by reduced sensory nerve and compound muscle action potentials. Neuropathy is typically detected in patients on chronic therapy, although it has been documented in those taking large doses for acute infections.
There are few reports on metronidazole-induced encephalopathy (222)(223).
Testosterone serum level is decreased by both high and therapeutic doses of metronidazole (224)(225)(226). In users of OCs less reduction of testosterone (by low dose estradiol OCs) is associated with less LPV-like signs and symptoms (178).
PKC, TRPV1 and the interaction between the chromatin remodeler BRG1, neuregulin-1, nuclear factor-kappaB (NF-κB) and the GR are common factors for cancer and demyelination. Recently it was found that a physical interaction in the cytosol is part of the GR- NF-κB cross-talk. (51)(78)(227)(228)(229)(230)(231)(232)(233)(234).
In conclusion/discussion: the hormone distortion, neurological complication, genotoxic, epigenetic and likely carcinogenic effects of metronidazole makes metronidazole a very likely cause of demyelination/LPV. In number of cases and rate of diagnosis, LPV is best described as the bulk of the iceberg. The underlying mechanism is likely to be malfunctions of GR-metronidazole replacing GR-cortisol functions.

Genes --> Stress, anxiety, depression
Genes associated with major depression and posttraumatic stress disorder have recently been identified (235)(236)(237)(238). A subtype of the corticotropin-releasing hormone receptor, CRHR1, has a key role in anxiety, depressive disorders and stress-associated pathologies (239).

Genes -->  Interleukin-1 --> RVVI’s
There is a genetic profile of women suffering of vulvodynia, especially genetic polymorphisms from genes coding for cytokines, IL-1 receptor antagonist (IL-1RA) and IL-1 β, and a gene coding for mannose-binding lectin (31).
The IL-1RA antagonist is a naturally occurring down-regulator of proinflammatory immune responses. In the gene coding for it, allele 2 was found homozygous in 52.9% of women with LPV, and only in 8.5 % of control women (240).
The gene coding for the IL-1RA is located on chromosome 2 in close proximity to the genes coding for IL-1 α  and IL-1β. IL-1α and IL-1β are major inducers of proinflammatory immune responses. IL-1RA is normally present in the circulation of healthy persons, whereas IL-1α and IL-1 β are not typically detectable in the absence of disease or autoimmunity. In some studies where persons homozygous for allele 2 of IL-1RA had higher circulating IL-1RA levels than did persons with other genotypes, IL-1β levels were also elevated. This resulted in the lowest IL-1RA/ IL-1β ratio and was associated with a heightened and prolonged proinflammatory immune response (241).
LPV patients with these genetic polymorphisms have a chronic unspecific inflammation and an inadequate inflammatory response, both in normal state and under infection (31).
This results in a greater frequency of candida and human papillomavirus infections and a higher frequency of allergy (242).  As mentioned in the introduction, LPV is also associated with a greater frequency of other infections. Whether these associations in part can be ascribed to genetic polymorphisms remains to be found out.

Interleukin 1 --> GR-cortisol --> kinases --> neural hyperplasia
IL-1β increases central melanocortin signalling by activating a subpopulation of proopiomelanocortin neurons in the arcuate nucleus of the hypothalamus and stimulating their release of melanocyte-stimulating hormone (243). Both IL-1α and IL-1β increase cortisol, androstenedione, dehydroepiandrosterone and dehydroepiandrosterone sulfate production in a human adrenocortical cell line (244). Genetic variations in the IL-1β gene contribute to the HPA axis alteration assessed by dexamethasone-suppression- test-cortisol in healthy subjects (245).
IL-1α induces activation of p38 mitogen-activated protein kinase and inhibits GR function. (246).
Corticosteroids can also be locally synthesized in various other tissues via locally expressed mediators of the hypothalamic-pituitary-adrenal (HPA) axis or renin-angiotensin system (RAS). . Local synthesis creates high corticosteroid concentrations in extra-adrenal organs, sometimes much higher than circulating concentrations. Locally synthesized GCs regulate activation of immune cells, while locally synthesized mineralocorticoids regulate blood volume and pressure (247). Both IL-1 and ACTH (adrenocorticotropic hormone) induce local cortisol synthesis in epidermis (248).
In melanocytes CRH (corticotropin-releasing hormone) stimulation of corticosteroids production is mediated by ACTH. The melanocyte response to CRH is highly organized along the same functional hierarchy as the HPA axis. This pattern demonstrates the fractal nature of the response to stress with similar activation sequence at the single-cell and whole body levels (249).
Melanocytes and SCs are derived from the multipotent population of neural crest cells,
and they are intimately interconnected far beyond previously postulated limits in that they share a common post-neural crest progenitor, i.e. the SC precursor (250). (Remember that SCs shifts to a progenitor-like state under demyelination and neural hyperplasia).
Long before any of the above was known, in 1995, Dyer et al. found: Binding of the stable melanocortin analogue Org2766 to cultured rat sciatic nerve SCs increased NGF receptors on SCs and evoked the release of neurotrophic factor(s) that synergized with NGF in stimulating neurite outgrowth. Thus SCs are a primary target for the action of melanocortins and melanocortins might stimulate neurite sprouting (251).
In a novel animal model of pruritus, induced by successive topical application of GC to mouse skin, NGF mRNA was slightly increased and remained high even after GC discontinuation. (252). GITR activation is required for the phosphorylation of ERK1 and ERK2 by NGF that is necessary for neurite growth (253).

Genes --> MBL --> GR-cortisol --> kinases --> demyelination --> pain
A single nucleotide polymorphism at codon 54 in the Mannose binding lectin (MBL) gene is associated with the development of primary LPV and a reduced capacity for TNF-α  production in response to microbial components. The MBL gene results in formation of an unstable MBL that is rapidly degraded. Thus, individuals carrying this MBL polymorphism have lowered circulating and vaginal MBL levels and they are more susceptible to a variety of infections (254).
MBL is an early complement factor that tag for innate immune recognition, which is needed for the inhibition of the primary MAPKs (ERK1/2, JNK, and particularly p38 MAPK) by naturally arising IgM antibodies. Such naturally arising IgM antibodies can suppress proinflammatory responses to purified agonists for Toll-like receptors (TLRs). The suppression of TLR-mediated MAPK signalling, correlates with, and has an absolute requirement for, the induction and nuclear localization of MAPK phosphatase-1, a prototypic counter-regulatory factor for the primary MAPKs known to mediate GC suppression of immune responses (255)
Lack of MBL thus results in insufficient stimulation of this particular GR-cortisol-controlled inhibitory pathway that can dampen pathogenic inflammatory responses. The lacking suppression of the primary MAPKs results in demyelination and mechanical allodynia (47)(108).

Recurrent vulvovaginal infections (RVVI’s) --> GR-cortisol
Bodily insults, including inflammation, pain, infection or even mental stress, lead to activation of the hypothalamic-pituitary-adrenal (HPA) axis, which stimulates the adrenal cortex to release GCs such as cortisol (256).
Farmer et al. showed that recurrent yeast infection in the mouse replicates important features of human provoked vulvodynia: mechanical allodynia and hyperinnervation localized to the vulva. Mechanical hypersensitivity persisted long after the resolution of the active infection. Long-lasting behavioural allodynia in a subset of mice was also observed after a single, extended Candida infection, as well as after repeated vulvar inflammation induced with zymosan, a mixture of fungal antigens. Only a subset of the infected mice exhibited LPV-like signs, which may indicate the importance of genetic background for the development of LPV. The other results indicate that LPV may be connected to increased immune response. The infected mice were in between infections treated with fluconazole, as were the placebo mice (28). As fluconazole does not interfere with the GR, fluconazole was an excellent choice.
Transient pre-treatment of healthy humans with cortisol induces a delayed systemic inflammatory response. This inflammatory response is maximal at an intermediate concentration, which approximates that observed in vivo following a major systemic inflammatory stimulus (257).
Both synthetic and endogenous GCs down-regulate GR mRNA level (258).
Increased expression of GRβ, which has a dominant negative effect on GRα-induced transactivation of GRE-driven promoters, could also be a possible underlying effect. However, GRβ has also intrinsic gene-specific transcriptional activity distinct from that of GRα (259)  Higher ratios of the expression level of hGRα/ hGRβ correlate with GC sensitivity, while lower ratios correlate with GC resistance (260).

SCs and the immune response
SCs express a plethora of pattern recognition receptors that allows them to recognize exogenous as well as endogenous danger signals. SCs initiate and regulate local immune responses by presenting antigens and by secreting pro- and anti-inflammatory cytokines, chemokines and neurotrophic factors, which will further attract immune cells. SCs express high levels of TLRs. By interacting with immune cells SCs contribute in shaping immune responses that can lead to inflammatory neuropathies (261)(262)(263)(264)(265).

Mast cells and SCs
The number of mast cells is increased in tender sites of the vulvar vestibule in LPV-patients compared with nontender sites and sites in control-women (29). Mast cell degranulation is accompanied by hyperalgesia, tissue edema, and neutrophils influx in the hindpaws of mice (266).
The SGK1 participates in the stimulation of Ca(2+) entry into and degranulation of mast cells (267).  Degranulation of mast cells is accompanied by release of heparanase, heparin, histamine and serotonin. SCs express both histamine and serotonin receptors (268)(269)(270)(271). Heparin is strongly alkaline. Alkaline pH causes pain sensation through activation of TRPA1 (272)(273). Mast cell-derived proteases can degrade myelin proteins, and myelin proteins or their breakdown products can potentiate further mast cell degranulation (274). Myelin debris is an important variable in the inflammatory response during demyelinating events (275)(276). The interactions of SC demyelination and mast cell degranulation may thus be highly relevant in the understanding of LPV. Further research is needed.

IV. LPV, GR-cortisol and the brain

Demyelination
Young women with relatively short-standing LPV (1 to 9 yrs) have, compared to controls, significantly higher grey (unmyelinated) matter densities in pain modulatory and stress-related areas of hippocampus and basal ganglia, which is related to lowered pain thresholds and increased pain catastrophizing scores (277).
Glutamate is released by stress and GC’s. (278)(279)(280). By a nongenomic action, GC enhances NMDA neurotoxicity through facilitating intracellular free calcium increment and attenuating the ERK1/2-mediated neuroprotective signalling (in a hippocampal neuron culture) (281).
Overactivation of ionotropic glutamate receptors in oligodendrocytes induces cytosolic Ca(2+) overload and excitotoxic death and demyelination. Intracellular Ca2+ release through ryanodine receptors contributes to this. In the white (myelinated) matter Ca(2+) influx into myelin induces myelin degradation in tissues and in vivo. Glutamate application results in paranodal myelin splitting and retraction. The break of axo-glial junctions exposes juxtaparanodal K+ channels, resulting in axonal conduction deficit (282)(283)(284).

GR-cortisol --> Neurotransmitters --> Sensitization, Hypersensibility
The GR-cortisol induced increased ERK and NMDA receptor activation is involved in stress-enhanced allodynia and enhanced central sensitization. The adult hippocampus remains sensitive to even brief exposures to cortisol (285)(286).
The central nucleus of the amygdala, CeA, has been identified as a site of nociceptive processing that is important for sensitization induced by peripheral injury. The metabotropic glutamate receptor 5 (mGluR5) is an integral component of nociceptive processing in the CeA. Pharmacological activation of mGluRs in the CeA of mice is sufficient to induce peripheral hypersensitivity in the absence of injury (287).

GR-cortisol --> Ionchannels
GR activation affects various properties of voltage-dependent Na+ and Ca2+ conductances in hippocampal CA1 neurons and in the basolateral amygdala, and thereby neuronal excitability in the cells of these areas of the brain (288)(289).

LPV (chronic pain) --> stress, anxiety, depression
More women with LPV show blunted morning awakening cortisol and report more signs of burnout, and emotional and bodily symptoms of stress compared with healthy control women of the same age (14). Vulvodynia increases the risk of both new and recurrent onset of depression and anxiety (15). Among treatment-seeking women with vulvodynia and with lifetime major depressive disorder, the majority (62.5%) reported that their first depressive episode occurred before the onset of vulvodynia (16).

GR-cortisol --> stress, anxiety, Depression
The balance between corticosteroid actions induced via activation of the MR and the GR determines the brain's response to stress. In addition to the delayed genomic role, membrane-associated nongenomic signalling of MRs and GRs play a major role for the coordination of a rapid adaptive response to stress (290)
Corticosteroids can exert maladaptive rather than adaptive effects when their actions via MRs and GRs are chronically unbalanced due to chronic stress (291).
GCs exert opposing rapid actions on glutamate and GABA release by activating divergent G protein signalling. The simultaneous rapid stimulation of nitric oxide and endocannabinoid synthesis by GCs has important implications for the impact of stress on the brain as well as on neural-immune interactions in the hypothalamus (292).
GRs are critical to the negative feedback process that inhibits additional GC release. Compared to males, female rats have fewer GRs and impaired GR translocation following chronic adolescent stress. Under conditions of chronic stress, attenuated negative feedback in females would result in hypercortisolemia, an endocrine state thought to cause depression. Sex differences in stress-related receptors thus shift females more easily into the development of mood and anxiety disorders (293).
GR function is impaired in major depression, resulting in reduced GR-mediated negative feedback (GC resistance) on the HPA axis. A lack of the 'positive' effects of cortisol on the brain, because of GC resistance, is likely to be involved in the pathogenesis of depression. (294)(295).
Chronic elevated levels of GCs down regulates the expression of mGluR5, and may thus contribute to impairments in glutamate neurotransmission in MDD (296).
Ineffective action of GC hormones on target tissues could lead to immune activation,  and GC resistance could be responsible for the enhanced vulnerability of depressed patients to develop neurodegenerative changes later in life (297).

Psychosocial environment --> Stress, anxiety, depression --> GR-cortisol
Young women exposed to an episodic stressor in the midst of chronic stress show increased cortisol output and reduced expression of GR mRNA. By contrast, when women has low levels of chronic stress, episodic events were associated with decreased cortisol output and increased GR mRNA Simultaneous exposure to episodic and chronic stress may create wear and tear on the body, whereas exposure to episodic stress in the context of a supportive environment may toughen the body, protecting it against subsequent stressors (298).
Strictly healthy caregivers of Alzheimer’s disease patients are significantly more stressed, anxious and depressed, but have similar cortisol levels, reduced DHEAS levels, increased cortisol/ DHEAS ratio, impaired HPA axis response to DEX intake, higher T cell proliferation and increased sensitivity to GCs compared to age-matched controls (299).

V. Conclusion and consequences

Conclusion: Although the evidence presented here might be considered circumstantial, the conviction must be:
Conclusion: LPV is caused by distortion of cortisol metabolism primarily in SCs, which causes demyelination and  increased immune response. Demyelination cause innervation of the epithelium and, via neurotransmitters and ionchannels, chronic pain. Interactions between the HPA-axis and the nerve- and immune systems make LPV a systemic condition also affecting the brain.
LPV is a mechanical allodynia. There is evidence, that demyelination cause mechanical allodynia. There is evidence of neural hyperplasia in LPV. There is evidence, that demyelination causes neural hyperplasia. There is evidence, that several GR-cortisol controlled factors cause both neural hyperplasia and demyelination. There is evidence, that LPV patients have experienced GR-cortisol distortions through the factors statistically associated with LPV.
LPV is caused by increased cortisol and/or decreased GR-function, simultaneously or consecutively. All the statistically suspected causes of LPV leads to increase of cortisol and decreased GR function. OCs increase cortisol through increase in CBG, and decrease GR function through ER-estradiol antagonism towards the GR. Repeated or major infections increase cortisol, which by it self leads to decreased GR expression. GCs, and some antifungals and antibiotics increase cortisol (/GC) level and reduce or inhibit GR function. While clotrimazole likely overactivates the GR, metronidazole likely malfunctions as a ligand for GR resulting in gentoxity and possibly cancer, and therefore both are likely causes of LPV.
The IL genetic polymorphism increase IL, which is associated with chronic elevated immune response, increased cortisol and decreased GR function. The MBL genetic polymorphism interrupts a specific immune response controlled by GR-cortisol, which via MAPKs cause in demyelination. Stress, anxiety and depression are also connected to high cortisol and/or impaired GR-function and demyelination.
LPV is caused by demyelination. Increased cortisol leads to increased expression of TRPV1, NMDA and increased intracellular calcium in glial cells, neuronal cells and in myelin itself, which leads to demyelination in the PNS and in stress related parts of the brain. Impaired GR function leads to lower GILZ, which affects the immune system, Na (+)-homeostasis and myelination. Demyelination and increased intracellular calcium makes sensory neuronal cells transmit pain signals instead of sensory signals.

Significance and consequences
Many million women now know why they suffer – provided the message is spread. This text is of course far from the full and final explanation of LPV, but it may be a beginning to that. Prevention, treatment of, and research in LPV can now take a direction and may advance in giant leaps for womankind.
The use of the different existing antifungals and antibiotics should be reconsidered in accordance with their hormone-disturbing effects. These effects should be fully clarified, and absence of such effects should be targeted in development and approval of new antifungals and antibiotics.
OCs, that do not have hormone-disturbing effects, are on the other hand at present hard to imagine. Nevertheless, these effects should be minimised, and these effects should be better known by research, doctors and users. Finally, the use of GCs should be reconsidered.
The research results presented here, and their combination, are not specific to LPV or vulvodynia, but may be relevant to many other illnesses: of course, those found co-morbid with LPV, but also other allodynias or even neuropathic pains in general, other diseases of mucous membranes, HPA-axis and myelin. – All the more reason to minimise the distortion of GR-cortisol (and other steroids and their receptors) caused by medicine. Use of alternatives to metronidazole and clotrimazole in the treatment of vulvovaginal infections would be a logical first step in this many-mile walk. 


Comments can be made and read below the references.


References

1. Haefner HK. Report of the International Society for the Study of Vulvovaginal Disease Terminology and Classification of Vulvodynia. Journal of Lower Genital Tract Disease, Issue: Volume 11(1), January 2007, pp 48-49.

2. Reed BD & Cantor LE. Vulvodynia in preadolescent girls. J Low Genit Tract Dis. 2008 Oct;12(4):257-61.

3. Selo-Ojeme DO, Paranjothy S & Onwude JL. Interstitial cystitis coexisting with vulvar vestibulitis in a 4-year-old girl. Int Urogynecol J Pelvic Floor Dysfunct. 2002;13(4):261-2.

4. Harlow BL, Wise LA and Stewart EG. Prevalence and predictors of chronic lower genital tract discomfort. Am J Obstet Gynecol. 2001 Sep;185(3):545-50.PMID: 11568775’

5. Edwards L (2004): Subsets of vulvodynia: overlapping characteristics. J Reprod Med. 2004 Nov;49(11):883-7.

6. Masheb RM et al. On the reliability and validity of physician ratings for vulvodynia and the discriminant validity of its subtypes. Pain Med. 2004 Dec; 5(4):349-58.

7. Bornstein J, Maman M & Abramovici H. "Primary" versus "secondary" vulvar vestibulitis: one disease, two variants. Am J Obstet Gynecol. 2001 Jan;184(2):28-31.

8. Wrosch C, Miller GE,  Schulz R. Cortisol Secretion and Functional Disabilities in Old Age: Importance of Using Adaptive Control Strategies. Psychosom Med 2009 November, 71(9): 996-1003.

9. Verdú E, Ceballos D, Vilches JJ, Navarro X. Influence of aging on peripheral nerve function and regeneration. J Peripher Nerv Syst. 2000 Dec;5(4):191-208.

10. Arnold LD, Bachmann GA, Rosen R, and Rhoads GG. Assessment of Vulvodynia Symptoms in a Sample of U.S. Women: A Prevalence Survey with a Nested Case Control Study. Am J Obstet Gynecol. 2007 February; 196(2): 128.e1–128.e6.

11. Reed BD, Harlow SD, Sen A, Legocki LJ, Edwards RM, et al. Prevalence and demographic characteristics of vulvodynia in a population-based sample. Am J Obstet Gynecol. 2011 Aug 22. [Epub ahead of print]

12. Kahn BS, Tatro C, Parsons CL and Willems JJ. Prevalence of interstitial cystitis in vulvodynia patients detected by bladder potassium sensitivity. J Sex Med. 2010 Feb;7(2 Pt 2):996-1002. Epub 2009 Oct 20.

13. Edgardh K and Abdelnoor M. Vulvar Vestibulitis and Risk Factors: a Population-based Casecontrol Study in Oslo. Acta Derm Venereol 2007; 87: 350–354. 

14. Ehrström S, Kornfeld D, Rylander E, Bohm-Starke N. Chronic stress in women with localised provoked vulvodynia. J Psychosom Obstet Gynaecol. 2009 Mar;30(1):73-9.

15. Khandker M, Brady SS, Vitonis AF, Maclehose RF, Stewart EG, Harlow BL. The influence of depression and anxiety on risk of adult onset vulvodynia. J Womens Health (Larchmt). 2011 Oct;20(10):1445-51. Epub 2011 Aug 8.

16. Masheb RM, Wang E, Lozano C and Kerns RD. Prevalence and correlates of depression in treatment-seeking women with vulvodynia. J Obstet Gynaecol. 2005 Nov;25(8):786-91.

17. Sjöberg I and Nylander Lundqvist EN. Vulvar vestibulitis in the north of Sweden. An epidemiologic case-control study. J Reprod Med. 1997 Mar;42(3):166-8.

18. Bouchard C, Brisson J, Fortier M, Morin C and Blanchette C (2002): Use of oral contraceptive pills and vulvar vestibulitis: a case-control study. Am J Epidemiol. 2002 Aug 1;156(3):254-61.

19. Berglund AL, Nigaard L and Rylander E. Vulvar pain, sexual behavior and genital infections in a young population: a pilot study. Acta Obstet Gynecol Scand. 2002 Aug; 81(8):738-42.

20. Bohm-Starke N, Johannesson U, Hilliges M, Rylander E and Torebjörk E. Decreased mechanical pain threshold in the vestibular mucosa of women using oral contraceptives: a contributing factor in vulvar vestibulitis? J Reprod Med. 2004 Nov;49(11):888-92.

21. Greenstein A, Ben-Aroya Z, Fass O, Militscher I, Roslik Y, Chen J and Abramov L. Vulvar vestibulitis syndrome and estrogen dose of oral contraceptive pills. J Sex Med. 2007 Nov;4(6):1679-83.

22. Goldstein A, Burrows L and Goldstein I. Can Oral Contraceptives Cause Vestibulodynia? J Sex Med 2010;7:1585–1587

23. Bohm-Starke N, Hilliges M, Falconer C and Rylander E (1998). Increased intraepithelial innervation in women with vulvar vestibulitis syndrome. Gynecol Obstet Invest. 1998; 46(4):256-60.

24. Weström LV & Willén R. Vestibular nerve fiber proliferation in vulvar vestibulitis syndrome. Obstet Gynecol. 1998 Apr;91(4):572-6.

25. Tympanidis P, Terenghi G and Dowd P- Increased innervation of the vulval vestibule in patients with vulvodynia. Br J Dermatol. 2003 May;148(5):1021-7.

26. Bornstein J, Goldschmid N and Sabo E. Hyperinnervation and mast cell activation may be used as histopathologic diagnostic criteria for vulvar vestibulitis. Gynecol Obstet Invest. 2004;58(3):171-8. Epub 2004 Jul 9.

27. Bohm-Starke N, Falconer C, Rylander E and Hilliges M. The expression of cyclooxygenase 2 and inducible nitric oxide synthase indicates no active inflammation in vulvar vestibulitis. Acta Obstet Gynecol Scand. 2001 Jul;80(7):638-44.)

28. Farmer MA, Taylor AM, Bailey AL et al. Repeated vulvovaginal fungal infections cause persistent pain in a mouse model of vulvodynia. Sci Transl Med. 2011 Sep 21;3(101):101ra91.

29. Goetsch MF, Morgan TK, Korcheva VB, Li H, Peters D, Leclair CM. Histologic and receptor analysis of primary and secondary vestibulodynia and controls: a prospective study. Am J Obstet Gynecol. 2010 Jun;202(6):614.e1-8. Epub 2010 Apr 28.

30. Harlow BL, He W, Nguyen RH. Allergic reactions and risk of vulvodynia. Ann Epidemiol. 2009 Nov;19(11):771-7.

31. Gerber S, Witkin SS and Stucki D. Immunological and genetic characterization of women with vulvodynia. J Med Life. 2008 Oct-Dec;1(4):432-8.

32. Bohm-Starke N, Hilliges M, Blomgren B, Falconer C and Rylander. Increased blood flow and erythema in the posterior vestibular mucosa in vulvar vestibulitis. Obstet Gynecol. 2001 Dec;98(6):1067-74.

33. Bohm-Starke N. Medical and physical predictors of localized provoked vulvodynia. Acta Obstet Gynecol Scand. 2010 Dec;89(12):1504-10.

34. Zhang Z, Zolnoun DA, Francisco EM, Holden JK, Dennis RG, Tommerdahl M. Altered central sensitization in subgroups of women with vulvodynia. Clin J Pain. 2011 Nov-Dec;27(9):755-63.

35. Gaitonde P, Rostron J, Longman L, Field EA. Burning mouth syndrome and vulvodynia coexisting in the same patient: a case report. Dent Update. 2002 Mar;29(2):75-6.

36. Feldhaus-Dahir M. The causes and prevalence of vestibulodynia: a vulvar pain disorder. Urol Nurs. 2011 Jan-Feb;31(1):51-4.

37.Wang Y, Khaing ZZ, Li N, Hall B, Schmidt CE and Ellington AD. Aptamer antagonists of myelin-derived inhibitors promote axon growth. PLoS One. 2010 Mar 16;5(3):e9726.

38. Wong EV, David S, Jacob MH, Jay DG. Inactivation of myelin-associated glycoprotein enhances optic nerve regeneration. J Neurosci. 2003 Apr 15;23(8):3112-7.

39. Kosins AM, et al. Improvement of Peripheral Nerve Regeneration Following Immunological Demyelination in Vivo. Plast Reconstr Surg. 2011 Jan 11. [Epub ahead of print]

40. Naik R. Neural hyperplasia in appendix. Indian journal of medical sciences. 01/1997; 50(12):339-41.

41. Güller U, Oertli D, Terracciano L, Harder F. [Neurogenic appendicopathy: a frequent, almost unknown disease picture. Evaluation of 816 appendices and review of the literature]. [Article in German]  Chirurg. 2001 Jun;72(6):684-9.

42. Arroyo EJ, Sirkowski EE, Chitale R, Scherer SS. Acute demyelination disrupts the molecular organization of peripheral nervous system nodes. J Comp Neurol. 2004 Nov 22;479(4):424-34.

43. Devor M: Sodium channels and mechanisms of neuropathic pain. J Pain. 2006 Jan;7(1 Suppl 1):S3-S12.

44. Boiko T, Rasband MN, Levinson SR, Caldwell JH, Mandel G, Trimmer JS, Matthews G. Compact myelin dictates the differential targeting of two sodium channel isoforms in the same axon. Neuron. 2001 Apr;30(1):91-104.

45. Matthew J. Craner, Albert C. Lo, Joel A. Black and Stephen G. Waxman Abnormal sodium channel distribution in optic nerve axons in a model of inflammatory demyelination. Brain (2003), 126, 1552-1561.

46. Black JA, Waxman SG, Smith KJ. Remyelination of dorsal column axons by endogenous Schwann cells restores the normal pattern of Nav1.6 and Kv1.2 at nodes of Ranvier. Brain. 2006 May;129(Pt 5):1319-29. Epub 2006 Mar 14.

47. Bhangoo S, Ren D, Miller RJ, Henry KJ, Lineswala J, et al. Delayed functional expression of neuronal chemokine receptors following focal nerve demyelination in the rat: a mechanism for the development of chronic sensitization of peripheral nociceptors. Mol Pain. 2007 Dec 12;3:38.

48. Okutsu M, Ishii K, Niu KJ, Nagatomi R. Cortisol-induced CXCR4 augmentation mobilizes T lymphocytes after acute physical stress. Am J Physiol Regul Integr Comp Physiol. 2005 Mar;288(3):R591-9. Epub 2004 Nov 4.

49. Okutsu M, Suzuki K, Ishijima T, Peake J, Higuchi M. The effects of acute exercise-induced cortisol on CCR2 expression on human monocytes. Brain Behav Immun. 2008 Oct;22(7):1066-71. Epub 2008 May 13.

50. Fernandes ES, Fernandes MA, Keeble JE. The functions of TRPA1 and TRPV1: moving away from sensory nerves. Br J Pharmacol. 2012 May;166(2):510-21. doi: 10.1111/j.1476-5381.2012.01851.x.

51. Tympanidis P, Casula MA, Yiangou Y, Terenghi G, Dowd P, Anand P. Increased vanilloid receptor VR1 innervation in vulvodynia. Eur J Pain. 2004 Apr;8(2):129-33.

52. Medvedeva YV, Kim MS, Usachev YM. Mechanisms of prolonged presynaptic Ca2+ signaling and glutamate release induced by TRPV1 activation in rat sensory neurons. J Neurosci. 2008 May 14;28(20):5295-311. doi: 10.1523/JNEUROSCI.4810-07.2008.

53. Lee J, Chung MK, Ro JY. Activation of NMDA receptors leads to phosphorylation of TRPV1 S800 by protein kinase C and A-Kinase anchoring protein 150 in rat trigeminal ganglia. Biochem Biophys Res Commun. 2012 Jul 27;424(2):358-63. doi: 10.1016/j.bbrc.2012.07.008. Epub 2012 Jul 10.

54. Smith KJ, Hall SM. Peripheral demyelination and remyelination initiated by the calcium-selective ionophore ionomycin: in vivo observations. J Neurol Sci. 1988 Jan;83(1):37-53.

55. Hogan QH. Role of Decreased Sensory Neuron Membrane Calcium Currents in the Genesis of    Neuropathic Pain. Croat Med J 2007;48:9-2.

56. Chevalier M, Gilbert G, Lory P, Marthan R, Quignard JF, Savineau JP. Dehydroepiandrosterone (DHEA) inhibits voltage-gated T-type calcium channels. Biochem Pharmacol. 2012 Jun 1;83(11):1530-9. doi: 10.1016/j.bcp.2012.02.025. Epub 2012 Mar 3.

57. Oloyo AK, Sofola OA, Nair RR, Harikrishnan VS, Fernandez AC. Testosterone relaxes abdominal aorta in male Sprague-Dawley rats by opening potassium (K(+)) channel and blockade of calcium (Ca(2+)) channel. Pathophysiology. 2011 Jun;18(3):247-53. Epub 2011 Mar 24.

58. Luoma JI, Kelley BG, Mermelstein PG. Progesterone inhibition of voltage-gated calcium channels is a potential neuroprotective mechanism against excitotoxicity. Steroids. 2011 Aug;76(9):845-55. Epub 2011 Mar 1.

59. Kelley BG, Mermelstein PG. Progesterone blocks multiple routes of ion flux. Mol Cell Neurosci. 2011 Oct;48(2):137-41. Epub 2011 Jul 19.

60. White JP, Urban L, Nagy I. TRPV1 function in health and disease. Curr Pharm Biotechnol. 2011 Jan 1;12(1):130-44.

61. Harlow BL, Stewart EG. A population-based assessment of chronic unexplained vulvar pain: have we underestimated the prevalence of vulvodynia? J Am Med Womens Assoc. 2003 Spring;58(2):82-8.

62. O'Neill J, Brock C, Olesen AE, Andresen T, Nilsson M, Dickenson AH. Unravelling the mystery of capsaicin: a tool to understand and treat pain. Pharmacol Rev. 2012 Oct;64(4):939-71. doi: 10.1124/pr.112.006163.

63. Vyklický L, Nováková-Tousová K, Benedikt J, Samad A, Touska F, Vlachová V. Calcium-dependent desensitization of vanilloid receptor TRPV1: a mechanism possibly involved in analgesia induced by topical application of capsaicin. Physiol Res. 2008;57 Suppl 3:S59-68. Epub 2008 May 13.

64. Chung MK, Jung SJ, Oh SB. Role of TRP channels in pain sensation. Adv Exp Med Biol. 2011;704:615-36. doi: 10.1007/978-94-007-0265-3_33.

65. Takeda M,  Tsuboi Y,  Kitagawa J,  Nakagawa K, Iwata K,  Matsumoto S. Potassium channels as a potential therapeutic target for trigeminal neuropathic and inflammatory pain. Molecular Pain 2011, 7:5 http://www.molecularpain.com/content/7/1/s5.

66. Tulleuda A, Cokic B, Callejo G, Saiani B, Serra J and Gasull X. TRESK channel contribution to nociceptive sensory neurons excitability: modulation by nerve injury. Molecular Pain 2011, 7:30
http://www.molecularpain.com/content/7/1/30.

67. Pexton T, Moeller-Bertram T, Schilling JM, Wallace MS. Targeting voltage-gated calcium channels for the treatment of neuropathic pain: a review of drug development. Expert Opin Investig Drugs. 2011 Sep;20(9):1277-84. Epub 2011 Jul 11.

68. Hong S, Zheng G, Wu X, Snider NT, Owyang C, Wiley JW. Corticosterone mediates reciprocal changes in CB 1 and TRPV1 receptors in primary sensory neurons in the chronically stressed rat. astroenterology. 2011 Feb;140(2):627-637.e4. doi: 10.1053/j.gastro.2010.11.003. Epub 2010 Nov 9.

69. Heise N, Shumilina E, Nurbaeva MK, Schmid E, Szteyn K, et al. Effect of dexamethasone on Na+/Ca2+ exchanger in dendritic cells. Am J Physiol Cell Physiol. 2011 Jun;300(6):C1306-13. doi: 10.1152/ajpcell.00396.2010. Epub 2011 Feb 9.

70. Derfoul A, Robertson NM, Lingrel JB, Hall DJ, Litwack G. Regulation of the human Na/K-ATPase beta1 gene promoter by mineralocorticoid and glucocorticoid receptors. J Biol Chem. 1998 Aug 14;273(33):20702-11.

71. Yun CC. Concerted roles of SGK1 and the Na+/H+ exchanger regulatory factor 2 (NHERF2) in regulation of NHE3. Cell Physiol Biochem. 2003;13(1):29-40.

72. Yun CC, Chen Y, Lang F. Glucocorticoid activation of Na(+)/H(+) exchanger isoform 3 revisited. The roles of SGK1 and NHERF2. J Biol Chem. 2002 Mar 8;277(10):7676-83. Epub 2001 Dec 21.

73. Bae YJ, Yoo JC, Park N, Kang D, Han J, Hwang E, Park JY, Hong SG. Acute Hypoxia Activates an ENaC-like Channel in Rat Pheochromocytoma (PC12) Cells. Korean J Physiol Pharmacol. 2013 Feb;17(1):57-64. doi: 10.4196/kjpp.2013.17.1.57. Epub 2013 Feb 14.

74. Itani OA, Auerbach SD, Husted RF, Volk KA, Ageloff S, et al. Glucocorticoid-stimulated lung epithelial Na(+) transport is associated with regulated ENaC and sgk1 expression. Am J Physiol Lung Cell Mol Physiol. 2002 Apr;282(4):L631-41.

75. Bergann T, Fromm A, Borden SA, Fromm M, Schulzke JD. Glucocorticoid receptor is indispensable for physiological responses to aldosterone in epithelial Na+ channel induction via the mineralocorticoid receptor in a human colonic cell line. Eur J Cell Biol. 2011 May;90(5):432-9. doi: 10.1016/j.ejcb.2011.01.001. Epub 2011 Feb 26.

76. Soundararajan R, Zhang TT, Wang J, Vandewalle A, Pearce D. A novel role for glucocorticoid-induced leucine zipper protein in epithelial sodium channel-mediated sodium transport. J Biol Chem. 2005 Dec 2;280(48):39970-81. Epub 2005 Oct 10.

77. Buse P, Maiyar AC, Failor KL, Tran S, Leong ML, Firestone GL. The stimulus-dependent co-localization of serum- and glucocorticoid-regulated protein kinase (Sgk) and Erk/MAPK in mammary tumor cells involves the mutual interaction with the importin-alpha nuclear import protein. Exp Cell Res. 2007 Sep 10;313(15):3261-75. Epub 2007 Jul 19.

78. Beck IM, Vanden Berghe W, Vermeulen L, Yamamoto KR, Haegeman G, De Bosscher K. Crosstalk in inflammation: the interplay of glucocorticoid receptor-based mechanisms and kinases and phosphatases. Endocr Rev. 2009 Dec;30(7):830-82. doi: 10.1210/er.2009-0013. Epub 2009 Nov 4.

79. Petersen CD, Giraldi A, Lundvall L, Kristensen E. Botulinum toxin type A-a novel treatment for provoked vestibulodynia? Results from a randomized, placebo controlled, double blinded study. J Sex Med. 2009 Sep;6(9):2523-37. doi: 10.1111/j.1743-6109.2009.01378.x. Epub 2009 Jul 10.

80. Traynelis SF, Wollmuth LP, McBain CJ, Menniti FS, Vance KM, et al. Glutamate receptor ion channels: structure, regulation, and function. Pharmacol Rev. 2010 Sep;62(3):405-96. doi: 10.1124/pr.109.002451.

81. Wang CC, Wang SJ. Modulation of presynaptic glucocorticoid receptors on glutamate release from rat hippocampal nerve terminals. Synapse. 2009 Sep;63(9):745-51.

82. Albert PR, Le François B and Millar AM. Transcriptional dysregulation of 5-HT1A
autoreceptors in mental illness. Molecular Brain 2011, 4:21.

83. Porter RJ, McAllister-Williams RH, Lunn BS, Young AH. 5-Hydroxytryptamine receptor function in humans is reduced by acute administration of hydrocortisone. Psychopharmacology (Berl). 1998 Oct;139(3):243-50.

84. Ou XM, Storring JM, Kushwaha N, Albert PR. Heterodimerization of mineralocorticoid and glucocorticoid receptors at a novel negative response element of the 5-HT1A receptor gene. J Biol Chem. 2001 Apr 27;276(17):14299-307. Epub 2001 Feb 2.

85. Chang LW, Viader A, Varghese N, Payton JE, Milbrandt J, Nagarajan R. An integrated approach to characterize transcription factor and microRNA regulatory networks involved in Schwann cell response to peripheral nerve injury. BMC Genomics. 2013 Feb 6;14:84. doi: 10.1186/1471-2164-14-84.

86. Svaren J, Meijer D. The molecular machinery of myelin gene transcription in Schwann cells. Glia. 2008 Nov 1;56(14):1541-51. doi: 10.1002/glia.20767.

87. Mittelstadt PR, Ashwell JD. Inhibition of AP-1 by the glucocorticoid-inducible protein GILZ. J Biol Chem. 2001 Aug 3;276(31):29603-10. Epub 2001 Jun 7.

88. Désarnaud F, Bidichandani S, Patel PI, Baulieu EE, Schumacher M. Glucocorticosteroids stimulate the activity of the promoters of peripheral myelin protein-22 and protein zero genes in Schwann cells. Brain Res. 2000 May 19;865(1):12-6.

89. Grenier J, Trousson A, Chauchereau A, Amazit L, Lamirand A, et al. Selective recruitment of p160 coactivators on glucocorticoid-regulated promoters in Schwann cells. Mol Endocrinol. 2004 Dec;18(12):2866-79. Epub 2004 Aug 26.

90. Fonte C, Grenier J, Trousson A, Chauchereau A, Lahuna O, et al. Involvement of {beta}-catenin and unusual behavior of CBP and p300 in glucocorticosteroid signaling in Schwann cells. Proc Natl Acad Sci U S A. 2005 Oct 4;102(40):14260-5. Epub 2005 Sep 26.

91. Newbern J, Birchmeier C. Nrg1/ErbB signaling networks in Schwann cell development and myelination. Semin Cell Dev Biol. 2010 Dec;21(9):922-8. doi: 10.1016/j.semcdb.2010.08.008. Epub 2010 Sep 9.

92. Syed N, Reddy K, Yang DP, Taveggia C, Salzer JL, Maurel P, Kim HA. Soluble neuregulin-1 has bifunctional, concentration-dependent effects on Schwann cell myelination. J Neurosci. 2010 Apr 28;30(17):6122-31. doi: 10.1523/JNEUROSCI.1681-09.2010.

93. Stassart RM, Fledrich R, Velanac V, Brinkmann BG, Schwab MH, et al. A role for Schwann cell-derived neuregulin-1 in remyelination. Nat Neurosci. 2013 Jan;16(1):48-54. doi: 10.1038/nn.3281. Epub 2012 Dec 9.

94. Chan JR, Phillips LJ 2nd and Glaser M. Glucocorticoids and progestins signal the initiation and enhance the rate of myelin formation. Proc Natl Acad Sci U S A. 1998 Sep 1;95(18):10459-64.

95. Abrahám IM, Meerlo P, Luiten PG. Concentration dependent actions of glucocorticoids on neuronal viability and survival. Dose-Response 2006: 4:38-54.

96. Lin W, Popko B. Endoplasmic reticulum stress in disorders of myelinating cells. Nat Neurosci. 2009 April; 12(4): 379–385.

97. D’Antonio M, Feltri ML and Wrabetz L. Myelin Under Stress. Journal of Neuroscience Research  (2009) 87:3241–3249.

98. Duret C, Daujat-Chavanieu M, Pascussi JM, Pichard-Garcia L, Balaguer P, Fabre JM, Vilarem MJ, Maurel P and Gerbal-Chaloin S. Ketoconazole and miconazole are antagonists of the human glucocorticoid receptor: consequences on the expression and function of the constitutive androstane receptor and the pregnane X receptor. Mol Pharmacol. 2006 Jul;70(1):329-39. Epub 2006 Apr 11.

99. Latasa MJ, Ituero M, Moran-Gonzalez A, Aranda A and Cosgaya JM. Retinoic acid regulates myelin formation in the peripheral nervous system. Glia. 2010 Sep;58(12):1451-64.

100. Puttagunta R, Schmandke A, Floriddia E, Gaub P, Fomin N, Ghyselinck NB, Di Giovanni S. RA-RAR-β counteracts myelin-dependent inhibition of neurite outgrowth via Lingo-1 repression. J Cell Biol. 2011 Jun 27;193(7):1147-56. doi: 10.1083/jcb.201102066. Epub 2011 Jun 20.

101. Lee X, Yang Z, Shao Z, Rosenberg SS, Levesque M, et al. NGF regulates the expression of axonal LINGO-1 to inhibit oligodendrocyte differentiation and myelination. J Neurosci. 2007 Jan 3;27(1):220-5.

102. Chrast R, Saher G , Nave K-A , Verheijen MHG. Lipid metabolism in myelinating glial cells: lessons from human inherited disorders and mouse models. Journal of Lipid Research Volume 52, 2011.

103. Yu CY, Mayba O, Lee JV, Tran J, Harris C, Speed TP, Wang JC. Genome-wide analysis of glucocorticoid receptor binding regions in adipocytes reveal gene network involved in triglyceride homeostasis. PLoS One. 2010 Dec 20;5(12):e15188.

104. Ji RR, Gereau RW 4th, Malcangio M, Strichartz GR. MAP kinase and pain. Brain Res Rev. 2009 Apr;60(1):135-48. doi: 10.1016/j.brainresrev.2008.12.011. Epub 2008 Dec 25.

105. Sarina, Yagi Y, Nakano O, Hashimoto T, Kimura K, et al. Induction of neurite outgrowth in PC12 cells by artemisinin through activation of ERK and p38 MAPK signaling pathways. Brain Res. 2012 Nov 1. pii: S0006-8993(12)01749-0. doi: 10.1016/j.brainres.2012.10.059. [Epub ahead of print]

106. Newbern JM, Snider WD. Bers-ERK Schwann cells coordinate nerve regeneration. Neuron. 2012 Feb 23;73(4):623-6.

107. Napoli I, Noon LA, Ribeiro S, Kerai AP, Parrinello S, et al. A central role for the ERK-signaling pathway in controlling Schwann cell plasticity and peripheral nerve regeneration in vivo. Neuron. 2012 Feb 23;73(4):729-42.

108. Yang DP, Kim J, Syed N, Tung YJ, Bhaskaran A, et al. p38 MAPK activation promotes denervated Schwann cell phenotype and functions as a negative regulator of Schwann cell differentiation and myelination. J Neurosci. 2012 May 23;32(21):7158-68. doi: 10.1523/JNEUROSCI.5812-11.2012.

109. Fragoso G, Robertson J, Athlan E, Tam E, Almazan G, Mushynski WE. Inhibition of p38 mitogen-activated protein kinase interferes with cell shape changes and gene expression associated with Schwann cell myelination. Exp Neurol. 2003 Sep;183(1):34-46.

110. Haines JD, Fragoso G, Hossain S, Mushynski WE, Almazan G. p38 Mitogen-activated protein kinase regulates myelination. J Mol Neurosci. 2008 May;35(1):23-33. Epub 2007 Nov 10.

111. Hossain S, de la Cruz-Morcillo MA, Sanchez-Prieto R, Almazan G. Mitogen-activated protein kinase p38 regulates Krox-20 to direct Schwann cell differentiation and peripheral myelination. Glia 2012 Jul;60(7):1130-44. doi: 10.1002/glia.22340. Epub 2012 Apr 17.

112. Haller J, Mikics E, Makara GB. The effects of non-genomic glucocorticoid mechanisms
on bodily functions and the central neural system. A critical evaluation of findings. Frontiers in Neuroendocrinology 29 (2008) 273–291.

113. Xu W, Shy M, Kamholz J, Elferink L, Xu G, Lilien J, Balsamo J. Mutations in the cytoplasmic domain of P0 reveal a role for PKC-mediated phosphorylation in adhesion and myelination. J Cell Biol. 2001 Oct 29;155(3):439-46. Epub 2001 Oct 22.

114. Guo L, Eviatar-Ribak T, Miskimins R. Sp1 phosphorylation is involved in myelin basic protein gene transcription. J Neurosci Res. 2010 Nov 15;88(15):3233-42. doi: 10.1002/jnr.22486.

115. Sivasankaran R, Pei J, Wang KC, Zhang YP, Shields CB, Xu XM, He Z. PKC mediates inhibitory effects of myelin and chondroitin sulfate proteoglycans on axonal regeneration. Nat Neurosci. 2004 Mar;7(3):261-8. Epub 2004 Feb 8.

116. Domeniconi M, Zampieri N, Spencer T, Hilaire M, Mellado W, Chao MV, Filbin MT. MAG induces regulated intramembrane proteolysis of the p75 neurotrophin receptor to inhibit neurite outgrowth. Neuron. 2005 Jun 16;46(6):849-55.

117. Maddali KK, Korzick DH, Turk JR, Bowles DK. Isoform-specific modulation of coronary artery PKC by glucocorticoids. Vascul Pharmacol. 2005 Mar;42(4):153-62.

118. Aziz MH, Shen H, Maki CG. Glucocorticoid receptor activation inhibits p53-induced apoptosis of MCF10Amyc cells via induction of protein kinase Cε. J Biol Chem. 2012 Aug 24;287(35):29825-36. doi: 10.1074/jbc.M112.393256. Epub 2012 Jul 6.

119. Aras-López R, Xavier FE, Ferrer M, Balfagón G. Dexamethasone decreases neuronal nitric oxide release in mesenteric arteries from hypertensive rats through decreased protein kinase C activation. Clin Sci (Lond). 2009 Aug 24;117(8):305-12. doi: 10.1042/CS20080178.

120. Lim G, Wang S, Zeng Q, Sung B, Yang L, Mao J. Expression of spinal NMDA receptor and PKCgamma after chronic morphine is regulated by spinal glucocorticoid receptor. J Neurosci. 2005 Nov 30;25(48):11145-54.

121. Micu I, Jiang Q, Coderre E, Ridsdale A, Zhang L, et al. NMDA receptors mediate calcium accumulation in myelin during chemical ischaemia. Nature. 2006 Feb 23;439(7079):988-92. Epub 2005 Dec 21.

122. Carozzi VA, Canta A, Oggioni N, Ceresa C, Marmiroli P et al. Expression and distribution of ‘high affinity’ glutamate transporters GLT1, GLAST, EAAC1 and of GCPII in the rat
peripheral nervous system. J. Anat. (2008) 213 1, pp539–546.

123. Perego C, Di Cairano ES, Ballabio M, Magnaghi V. Neurosteroid allopregnanolone regulates EAAC1-mediated glutamate uptake and triggers actin changes in Schwann cells. ).   J Cell Physiol. 2012 Apr;227(4):1740-51. doi: 10.1002/jcp.22898.

124. Suchak SK, Baloyianni NV, Perkinton MS, Williams RJ, Meldrum BS, Rattray M. The 'glial' glutamate transporter, EAAT2 (Glt-1) accounts for high affinity glutamate uptake into adult rodent nerve endings. J Neurochem. 2003 Feb;84(3):522-32.

125. Tao Z, Rosental N, Kanner BI, Gameiro A, Mwaura J, Grewer C. Mechanism of cation binding to the glutamate transporter EAAC1 probed with mutation of the conserved amino acid residue Thr101. J Biol Chem. 2010 Jun 4;285(23):17725-33. doi: 10.1074/jbc.M110.121798. Epub 2010 Apr 8.

126. García-Tardón N, González-González IM, Martínez-Villarreal J, Fernández-Sánchez E, Giménez C, Zafra F. Protein kinase C (PKC)-promoted endocytosis of glutamate transporter GLT-1 requires ubiquitin ligase Nedd4-2-dependent ubiquitination but not phosphorylation. J Biol Chem. 2012 Jun 1;287(23):19177-87. doi: 10.1074/jbc.M112.355909. Epub 2012 Apr 13.

127. Cremona ML, Matthies HJ, Pau K, Bowton E, Speed N, et al.. Flotillin-1 is essential for PKC-triggered endocytosis and membrane microdomain localization of DAT. Nat Neurosci. 2011 Apr;14(4):469-77. doi: 10.1038/nn.2781. Epub 2011 Mar 13.

128. Boehmer C, Palmada M, Rajamanickam J, Schniepp R, Amara S, Lang F. Post-translational regulation of EAAT2 function by co-expressed ubiquitin ligase Nedd4-2 is impacted by SGK kinases. J Neurochem. 2006 May;97(4):911-21. Epub 2006 Mar 29.

129. Guillet BA, Velly LJ, Canolle B, Masmejean FM, Nieoullon AL, Pisano P. Differential regulation by protein kinases of activity and cell surface expression of glutamate transporters in neuron-enriched cultures. Neurochem Int. 2005 Mar;46(4):337-46. Epub 2005 Jan 13.

130. Kuroda H, Sobhan U, Sato M, Tsumura M, Ichinohe T, Tazaki M, Shibukawa Y. Sodium-calcium exchangers in rat trigeminal ganglion neurons. Mol Pain. 2013 Apr 29;9(1):22. doi: 10.1186/1744-8069-9-22.

131. Boscia F, D'Avanzo C, Pannaccione A, Secondo A, Casamassa A, et al. Silencing or knocking out the Na(+)/Ca(2+) exchanger-3 (NCX3) impairs oligodendrocyte differentiation. Cell Death Differ. 2012 Apr;19(4):562-72. doi: 10.1038/cdd.2011.125. Epub 2011 Sep 30.

132. Sirabella R, Secondo A, Pannaccione A, Molinaro P, Formisano L, et al. ERK1/2, p38, and JNK regulate the expression and the activity of the three isoforms of the Na+ /Ca2+ exchanger, NCX1, NCX2, and NCX3, in neuronal PC12 cells. J Neurochem. 2012 Sep;122(5):911-22. doi: 10.1111/j.1471-4159.2012.07838.x. Epub 2012 Jul 11.

133. Formisano L, Saggese M, Secondo A, Sirabella R, Vito P, et al. The two isoforms of the Na+/Ca2+ exchanger, NCX1 and NCX3, constitute novel additional targets for the prosurvival action of Akt/protein kinase B pathway. Mol Pharmacol. 2008 Mar;73(3):727-37. Epub 2007 Dec 13.

134. Long Y, Wang WP, Yuan H, Ma SP, Feng N, Wang L, Wang XL. Functional comparison of the reverse mode of Na+/Ca2+ exchangers NCX1.1 and NCX1.5 expressed in CHO cells. Acta Pharmacol Sin. 2013 May;34(5):691-8. doi: 10.1038/aps.2013.4. Epub 2013 Apr 8.

135. Kiedrowski L, Czyz A, Baranauskas G, Li XF, Lytton J. Differential contribution of plasmalemmal Na/Ca exchange isoforms to sodium-dependent calcium influx and NMDA excitotoxicity in depolarized neurons. J Neurochem. 2004 Jul;90(1):117-28.

136. Magi S, Lariccia V, Castaldo P, Arcangeli S, Nasti AA, Giordano A, Amoroso S. Physical and functional interaction of NCX1 and EAAC1 transporters leading to glutamate-enhanced ATP production in brain mitochondria. PLoS One. 2012;7(3):e34015. doi: 10.1371/journal.pone.0034015. Epub 2012 Mar 30.

137. Nikolaeva MA, Mukherjee B, Stys PK. Na+-dependent sources of intra-axonal Ca2+ release in rat optic nerve during in vitro chemical ischemia. J Neurosci. 2005 Oct 26;25(43):9960-7.

138. Nobbio L, Sturla L, Fiorese F, Usai C, Basile G, et al. P2X7 mediated increased intracellular calcium causes functional derangement in Schwann cells from rats with CMT1A neuropathy. J Biol Chem. 2009 Aug 21;284(34):23146-58. doi: 10.1074/jbc.M109.027128. Epub 2009 Jun 22.

139. Nakajima C, Kulik A, Frotscher M, Herz J, Schäfer M, Bock HH, May P. LDL receptor-related protein 1 (LRP1) modulates N-methyl-D-aspartate (NMDA) receptor-dependent intracellular signaling and NMDA-induced regulation of postsynaptic protein complexes. J Biol Chem. 2013 Jun 11. [Epub ahead of print]

140. Orita S, Henry K, Mantuano E, Yamauchi K, De Corato A, et al. Schwann cell LRP1 regulates remak bundle ultrastructure and axonal interactions to prevent neuropathic pain. J Neurosci. 2013 Mar 27;33(13):5590-602. doi: 10.1523/JNEUROSCI.3342-12.2013.

141. Kancha RK, Hussain MM. Up-regulation of the low density lipoprotein receptor-related protein by dexamethasone in HepG2 cells. Biochim Biophys Acta. 1996 Jun 11;1301(3):213-20.

142. Nilsson A, Vesterlund L, Oldenborg PA. Macrophage expression of LRP1, a receptor for apoptotic cells and unopsonized erythrocytes, can be regulated by glucocorticoids. Biochem Biophys Res Commun. 2012 Jan 27;417(4):1304-9. doi: 10.1016/j.bbrc.2011.12.137. Epub 2012 Jan 3.

143. Mantuano E, Henry K, Yamauchi T, Hiramatsu N, Yamauchi K, et al. The unfolded protein response is a major mechanism by which LRP1 regulates Schwann cell survival after injury. J Neurosci. 2011 Sep 21;31(38):13376-85. doi: 10.1523/JNEUROSCI.2850-11.2011.

144. Borcherding DC, Hugo ER, Idelman G, De Silva A, Richtand NW, Loftus J and Ben-Jonathan N. Dopamine Receptors in Human Adipocytes: Expression and Functions. PLoS ONE. www.plosone.org.1 September 2011, Volume 6,  Issue 9, e25537.

145. Natarajan A, Han G, Chen S-y, Yu P, White R, Jose P. The D5 Dopamine Receptor Mediates Large-Conductance, Calcium- and Voltage-Activated Potassium Channel Activation in Human Coronary Artery Smooth Muscle Cells.  Journal of Pharmacology and Experimental Therapeutics, 2010, Vol. 332, No. 2, 640–649.

146. Eckhardt M, Hedayati KK, Pitsch J, Lüllmann-Rauch R, Beck H, Fewou SN, Gieselmann V: Sulfatide storage in neurons causes hyperexcitability and axonal degeneration in a mouse model of metachromatic leukodystrophy. J Neurosci. 2007 Aug 22;27(34):9009-21.

147. Stephens JL and Pieringer RA. Regulation of arylsulphatase A and sulphogalactolipid turnover by cortisol in myelinogenic cultures of cells dissociated from embryonic mouse brain. Biochem J. 1984 May 1;219(3):689-97.

148. Marcelo AJ, Pieringer RA. Hydrocortisone regulates arylsulfatase A (cerebroside-3-sulfate-3-sulfohydrolase) by decreasing the quantity of the enzyme in cultures of cells dissociated from embryonic mouse cerebra. Neurochem Res. 1990 Sep;15(9):937-44.

149. Adilakshmi T, Sudol I, Tapinos N. Combinatorial action of miRNAs regulates transcriptional and post-transcriptional gene silencing following in vivo PNS injury. PLoS One. 2012;7(7):e39674. doi: 10.1371/journal.pone.0039674. Epub 2012 Jul 6.

150. Viader A, Chang LW, Fahrner T, Nagarajan R, Milbrandt J. MicroRNAs modulate Schwann cell response to nerve injury by reinforcing transcriptional silencing of dedifferentiation-related genes. J Neurosci. 2011 Nov 30;31(48):17358-69. doi: 10.1523/JNEUROSCI.3931-11.2011.

151. Yu B, Zhou S, Wang Y, Qian T, Ding G, Ding F, Gu X. miR-221 and miR-222 promote Schwann cell proliferation and migration by targeting LASS2 after sciatic nerve injury. J Cell Sci. 2012 Jun 1;125(Pt 11):2675-83. doi: 10.1242/jcs.098996. Epub 2012 Mar 5.

152. Zhou S, Shen D, Wang Y, Gong L, Tang X, et al. microRNA-222 targeting PTEN promotes neurite outgrowth from adult dorsal root ganglion neurons following sciatic nerve transection. PLoS One. 2012;7(9):e44768. doi: 10.1371/journal.pone.0044768. Epub 2012 Sep 13.

153. Sims RJ 3rd, Mandal SS, Reinberg D. Recent highlights of RNA-polymerase-II-mediated transcription. Curr Opin Cell Biol. 2004 Jun;16(3):263-71.

154. Schanen BC, Li X. Transcriptional regulation of mammalian miRNA genes. Genomics. 2011 Jan;97(1):1-6. doi: 10.1016/j.ygeno.2010.10.005. Epub 2010 Oct 23.

155. McEwan IJ, Almlöf T, Wikström AC, Dahlman-Wright K, Wright AP, Gustafsson JA. The glucocorticoid receptor functions at multiple steps during transcription initiation by RNA polymerase II. J Biol Chem. 1994 Oct 14;269(41):25629-36.

156. Martini L, Magnaghi V and Melcangi RC. Actions of progesterone and its 5alpha-reduced metabolites on the major proteins of the myelin of the peripheral nervous system. Steroids. 2003 Nov;68(10-13):825-9.

157. Robaglia-Schlupp A, Pizant J, Norreel J-C, Passage E , SabeÂran-Djoneidi D. et al. PMP22 overexpression causes dysmyelination in Mice. Brain (2002), 125, 2213±2221

158. Robert F, Guennoun R, DeÂsarnaud F, Do-Thi A, Benmessahel Y, Baulieu EE, Schumacher M. Synthesis of progesterone in Schwann cells: regulation by sensory neurons. European Journal of Neuroscience, 2001, Vol. 13, pp. 916-924.

159. Magnaghi V, Ballabio M, Gonzalez LC, Leonelli E, Motta M, Melcangi RC. The synthesis of glycoprotein Po and peripheral myelin protein 22 in sciatic nerve of male rats is modulated by testosterone metabolites. Molecular Brain Research 126 (2004) 67–73.

160. Morris DJ, Latif SA, Rokaw MD, Watlington CO, Johnson AP. A second enzyme protecting mineralocorticoid receptors from glucocorticoid occupancy. Am J Physiol Cell Physiol 274:C1245-C1252, 1998.

161. Hennebert O, Chalbot S, Alran S, Morfin R. Dehydroepiandrosterone 7alpha-hydroxylation in human tissues: possible interference with type 1 11beta-hydroxysteroid dehydrogenase-mediated processes. J Steroid Biochem Mol Biol. 2007 May;104(3-5):326-33. Epub 2007 Mar 24.

162. Zhou HY, Hu GX, Lian QQ, Morris D, Ge RS. The metabolism of steroids, toxins and drugs by 11β-hydroxysteroid dehydrogenase 1. Toxicology. 2012 Feb 6;292(1):1-12. Epub 2011 Nov 28.

163. Groyer G, Eychenne B, Girard C, Rajkowski K, Schumacher M, Cadepond F. Expression and functional state of the corticosteroid receptors and 11 beta-hydroxysteroid dehydrogenase type 2 in Schwann cells. Endocrinology. 2006 Sep;147(9):4339-50. Epub 2006 Jun 8.

164. Oakley RH, Cidlowski JA. Cellular processing of the glucocorticoid receptor gene and protein: new mechanisms for generating tissue-specific actions of glucocorticoids. J Biol Chem. 2011 Feb 4;286(5):3177-84. doi: 10.1074/jbc.R110.179325. Epub 2010 Dec 13. 183.

165. Golikov PP. [Effect of antibiotics on glucocorticoid receptor function]. [Article in Russian] Antibiot Khimioter. 1995 Jul;40(7):25-9.

166. Rezaii T, Ernberg M. Influence of oral contraceptives on endogenous pain control in healthy women. Exp Brain Res (2010) 203:329–338.

167. Perogamvros I, Aarons L, Miller AG, Trainer PJ, Ray DW. Source Corticosteroid-binding globulin regulates cortisol pharmacokinetics. Clin Endocrinol (Oxf). 2011 Jan;74(1):30-6. doi: 10.1111/j.1365-2265.2010.03897.x.

168. Coenen CM, Thomas CM, Borm GF, Rolland R. Comparative evaluation of the androgenicity of four low-dose, fixed-combination oral contraceptives. Int J Fertil Menopausal Stud. 1995; 40 Suppl 2:92-7.

169. Cachrimanidou AC, Hellberg D, Nilsson S, von Schoulz B, Crona N, Siegbahn A. Hemostasis profile and lipid metabolism with long-interval use of a desogestrel-containing oral contraceptive. Contraception. 1994 Aug;50(2):153-65.

170. Wiegratz I, Kutschera E, Lee JH, Moore C, Mellinger U, Winkler UH, Kuhl H. Effect of four different oral contraceptives on various sex hormones and serum-binding globulins. Contraception. 2003 Jan;67(1):25-32.

171. Qureshi AC, Bahri A, Breen LA, Barnes SC, Powrie JK, Thomas SM, Carroll PV. The influence of the route of oestrogen administration on serum levels of cortisol-binding globulin and total cortisol. Clin Endocrinol (Oxf). 2007 May;66(5):632-5.

172. Agren UM, Anttila M, Mäenpää-Liukko K, Rantala ML, Rautiainen H, Sommer WF, Mommers E. Effects of a monophasic combined oral contraceptive containing nomegestrol acetate and 17β-oestradiol in comparison to one containing levonorgestrel and ethinylestradiol on markers of endocrine function. Eur J Contracept Reprod Health Care. 2011 Sep 26. [Epub ahead of print]

173. Henley DE, Lightman SL. New insights into corticosteroid-binding globulin and glucocorticoid delivery. Neuroscience 180 (2011) 1–8.

174. Petersen HH,  Andreassen TK,  Breiderhoff T, Bräsen JH, Schulz H, et al. Hyporesponsiveness to Glucocorticoids in Mice Genetically Deficient for the Corticosteroid Binding Globulin.Molecular and cellular biology, Oct. 2006 Vol. 26, No. 19, p. 7236–7245.

175. Richard EM, Helbling JC, Tridon C, Desmedt A, Minni AM, et al. Plasma transcortin influences endocrine and behavioral stress responses in mice. Endocrinology. 2010 Feb;151(2):649-59. Epub 2009 Dec 18.

176. Perogamvros I, Underhill C, Henley DE, Hadfield KD, Newman WG, et al. Novel corticosteroid-binding globulin variant that lacks steroid binding activity. J Clin Endocrinol Metab. 2010 Oct;95(10):E142-50. Epub 2010 Jul 7.

177. Gagliardia L, Hob JT, Torpya DJ. Corticosteroid-binding globulin: The clinical significance of altered levels and heritable mutations. Molecular and Cellular Endocrinology 316 (2010) 24–34

178. She Y. Are there differences? SHBG levels, free testosterone and sexual pain in women using combined hormonal contraception: A retrospective study in vulvodynia patients. Abstracts / Contraception 82 (2010) 183–216.

179. Varea O, Garrido JJ, Dopazo A, Mendez P, Garcia-Segura LM, Wandosell F. Estradiol activates beta-catenin dependent transcription in neurons. PLoS One. 2009;4(4):e5153. doi: 10.1371/journal.pone.0005153. Epub 2009 Apr 10.

180. Whirledge S, Cidlowski JA. Estradiol antagonism of glucocorticoid-induced GILZ expression in human uterine epithelial cells and murine uterus. Endocrinology. 2013 Jan;154(1):499-510. doi: 10.1210/en.2012-1748. Epub 2012 Nov 26.

181. Krishnan AV, Swami S, Feldman D. Estradiol inhibits glucocorticoid receptor expression and induces glucocorticoid resistance in MCF-7 human breast cancer cells. J Steroid Biochem Mol Biol. 2001 Apr;77(1):29-37.

182. Kinyamu HK, Archer TK. Estrogen receptor-dependent proteasomal degradation of the glucocorticoid receptor is coupled to an increase in mdm2 protein expression. Mol Cell Biol. 2003 Aug;23(16):5867-81.

183. Zhang Y, Leung DY, Nordeen SK, Goleva E. Estrogen inhibits glucocorticoid action via protein phosphatase 5 (PP5)-mediated glucocorticoid receptor dephosphorylation. J Biol Chem. 2009 Sep 4;284(36):24542-52. doi: 10.1074/jbc.M109.021469. Epub 2009 Jul 8.

184. Weiser MJ, Handa RJ. Estrogen impairs glucocorticoid dependent negative feedback on the hypothalamic-pituitary-adrenal axis via estrogen receptor alpha within the hypothalamus. Neuroscience. 2009 Mar 17;159(2):883-95. doi: 10.1016/j.neuroscience.2008.12.058. Epub 2009 Jan 7.

185. Loose DS, Stover EP, Feldman D. Ketoconazole Binds to Glucocorticoid Receptors
and Exhibits Glucocorticoid Antagonist Activity in Cultured Cells. J. Clin. Invest. Volume 72 July 1983 404-408.

186. Zhou SF. Drugs behave as substrates, inhibitors and inducers of human cytochrome P450 3A4. Curr Drug Metab. 2008 May;9(4):310-22.

187. Dvorak Z. Drug–drug interactions by azole antifungals: Beyond a dogma of CYP3A4 enzyme activity inhibition. Toxicology Letters 202 (2011) 129–132.

188. Takezawa T, Matsunaga T, Aikawa K, Nakamura K, Ohmori S. Lower expression of HNF4α and PGC1α might impair rifampicin-mediated CYP3A4 induction under conditions where PXR overexpressed in human fetal liver cells. Drug Metab Pharmacokinet. 2012 Feb 14. [Epub ahead of print]

189. Goetz AK, Dix DJ. Mode of action for reproductive and hepatic toxicity inferred from a genomic study of triazole antifungals. Toxicol Sci. 2009 Aug;110(2):449-62. Epub 2009 May 7.

190. Kjærstad MB, Taxvig C, Nellemann C, Vinggaard AM and Andersen HR. Endocrine disrupting effects in vitro of conazole antifungals used as pesticides and pharmaceuticals Reproductive Toxicology Volume 30, Issue 4, December 2010, Pages 573-582.

191. Meseguer V,  Karashima Y, Talavera K, D’Hoed D, Donovan-Rodrı´guez T, et al. Transient Receptor Potential Channels in Sensory Neurons Are Targets of the Antimycotic Agent Clotrimazole. The Journal of Neuroscience, January 16, 2008 • 28(3):576 –586.

192. Xi N, Bo Y, Doherty EM, Fotsch C, Gavva NR, et al. Synthesis and evaluation of thiazole carboxamides as vanilloid receptor 1 (TRPV1) antagonists. Bioorg Med Chem Lett. 2005 Dec 1;15(23):5211-7. Epub 2005 Oct 3.

193. Gore VK, Ma VV, Tamir R, Gavva NR, Treanor JJ, Norman MH. Structure-activity relationship (SAR) investigations of substituted imidazole analogs as TRPV1 antagonists. Bioorg Med Chem Lett. 2007 Nov 1;17(21):5825-30. Epub 2007 Aug 25.

194. Majerovich JA, Canty A, Miedema B. Chronic vulvar irritation: could toilet paper be the culprit? Canadian Family Physician Le Médecin de famille canadien Vol 56: april avril 2010.

195. Gravel A, Vijayan MM. Salicylate disrupts interrenal steroidogenesis and brain glucocorticoid receptor expression in rainbow trout. Toxicol Sci. 2006 Sep;93(1):41-9. Epub 2006 Mar 21.

196. Gong N, Zhang M, Zhang XB, Chen L, Sun GC, Xu TL. The aspirin metabolite salicylate enhances neuronal excitation in rat hippocampal CA1 area through reducing GABAergic inhibition. Neuropharmacology. 2008 Feb;54(2):454-63. Epub 2007 Nov 6.

197. Golikov PP, Nikolaeva NIu. [Effect of sodium salicylate on the function of glucocorticoid receptors type II and III]. [Article in Russian] Patol Fiziol Eksp Ter. 1995 Apr-Jun;(2):13-5.

198. Golikov PP, Nikolaeva NIu, Marchenko VV. [Nonsteroidal antiinflammatory agents as modulators of glucocorticoid function of receptors]. [Article in Russian] Vestn Ross Akad Med Nauk. 1994;(2):47-52.

199. Feldman D, Funder JW, Edelman IS. Evidence for a new class of corticosterone receptors in the rat kidney. Endocrinology. 1973 May;92(5):1429-41.

200. Náray-Fejes-Tóth A, Rusvai E, Fejes-Tóth G. Is the renal type III corticosteroid-binding site the collecting duct-specific isoform of 11 beta-hydroxysteroid dehydrogenase? Endocrinology. 1994 Apr;134(4):1671-5.

201. Kristensen DM, Hass U, Lesné L, Lottrup G, Jacobsen PR,  et al. Intrauterine exposure to mild analgesics is a risk factor for development of male reproductive disorders in human and rat. Hum Reprod. 2011 Jan;26(1):235-44. Epub 2010 Nov 8.

202. Yildiz HY, Altunay S. Physiological stress and innate immune response in gilthead sea bream (Sparus aurata) and sea bass (Dicentrarchus labrax) exposed to combination of trimethoprim and sulfamethoxazole (TMP-SMX). Fish Physiol Biochem. 2011 Sep;37(3):401-9. Epub 2010 Oct 6.

203. Gardella B, Porru D, Nappi RE, Daccò MD, Chiesa A and Spinillo A. Interstitial Cystitis is Associated with Vulvodynia and Sexual Dysfunction-A Case-Control Study. J Sex Med. 2011 Apr 7. doi: 10.1111/j.1743-6109.2011.02251.x. [Epub ahead of print]

204. Grill MF and Manganti RK. Neurotoxic effects associated with antibiotic use: management considerations. Br J Clin Pharmacol / 72:3 / 381–393.

205. Gao YJ, Zhang L, Samad OA, Suter MR, Yasuhiko K, et al. JNK-induced MCP-1 production in spinal cord astrocytes contributes to central sensitization and neuropathic pain. J Neurosci. 2009 Apr 1;29(13):4096-108.

206. Gao YJ and  Ji RR. Chemokines, neuronal-glial interactions, and central processing of neuropathic pain Pharmacol Ther. 2010 April; 126(1): 56–68.

207. Kawasaki Y, Zhang L, Cheng JK, Ji RR. Cytokine mechanisms of central sensitization: distinct and overlapping role of interleukin-1beta, interleukin-6, and tumor necrosis factor-alpha in regulating synaptic and neuronal activity in the superficial spinal cord. J Neurosci. 2008 May 14;28(20):5189-94.

208. Placke T, Kopp HG, Salih HR. Glucocorticoid-induced TNFR-related (GITR) protein and its ligand in antitumor immunity: functional role and therapeutic modulation. Clin Dev Immunol. 2010;2010:239083. Epub 2010 Sep 26.

209. United States Department of Health and Human Services, National Toxicology Program.Report on Carcinogens, Twelfth Edition 2011 p.269.  http://ntp.niehs.nih.gov/go/roc12

210. Stover KR, Riche DM, Gandy CL, Henderson H. What would we do without metronidazole? Am J Med Sci. 2012 Apr;343(4):316-9.

211. Gillet E, Meys JF, Verstraelen H, Verhelst R, De Sutter P et al. Association between bacterial vaginosis and cervical intraepithelial neoplasia: systematic review and meta-analysis. PLoS One. 2012;7(10):e45201. doi: 10.1371/journal.pone.0045201. Epub 2012 Oct 2.

212. Pan SY, Morrison H. Epidemiology of cancer of the small intestine. World J Gastrointest Oncol. 2011 Mar 15;3(3):33-42. doi: 10.4251/wjgo.v3.i3.33.

213. Herszenyi L, Miheller P, Tulassay Z. Carcinogenesis in inflammatory bowel disease. Dig Dis. 2007;25(3):267-9.

214. López Nigro MM, Palermo AM, Mudry MD, Carballo MA. Source Cytogenetic evaluation of two nitroimidazole derivatives. Toxicol In Vitro. 2003 Feb;17(1):35-40.

215. Mudry MD, Martinez RA, Nieves M, Carballo MA. Mutat Res. Biomarkers of genotoxicity and genomic instability in a non-human primate, Cebus libidinosus (Cebidae, Platyrrhini), exposed to nitroimidazole derivatives. 2011 Mar 18;721(1):108-13. doi: 10.1016/j.mrgentox.2011.01.002. Epub 2011 Jan 19.

216. Ana Carballo M, Martinez RA, Mudry MD. Nitroimidazole derivatives: non-randomness sister chromatid exchanges in human peripheral blood lymphocytes. J Appl Toxicol. 2009 Apr;29(3):248-54. doi: 10.1002/jat.1403.

217. Gómez-Arroyo S, Melchor-Castro S, Villalobos-Pietrini R, Camargo EM, Salgado-Zamora H, Campos Aldrete ME. Cytogenetic study of metronidazole and three metronidazole analogues in cultured human lymphocytes with and without metabolic activation. Toxicol In Vitro. 2004 Jun;18(3):319-24.

218. Deroo BJ, Archer TK. Glucocorticoid receptor-mediated chromatin remodeling in vivo. Oncogene. 2001 May 28;20(24):3039-46.

219. Zhang L, Li H, Hu X, Li XX, Smerin S, Ursano R. Glucocorticoid-induced p11 over-expression and chromatin remodeling: a novel molecular mechanism of traumatic stress? Med Hypotheses. 2011 Jun;76(6):774-7. doi: 10.1016/j.mehy.2011.02.015. Epub 2011 Mar 1.

220. Vicent GP, Zaurin R, Ballaré C, Nacht AS, Beato M.  Erk signaling and chromatin remodeling in MMTV promoter activation by progestins. Nuclear Receptor Signaling (2009) 7, e008.

221. Koturbash I, Beland FA, Pogribny IP. Role of epigenetic events in chemical carcinogenesis--a justification for incorporating epigenetic evaluations in cancer risk assessment. Toxicol Mech Methods. 2011 May;21(4):289-97. doi: 10.3109/15376516.2011.557881.

222. Hobson-Webb LD, Roach ES, Donofrio PD. Metronidazole: newly recognized cause of autonomic neuropathy. J Child Neurol. 2006 May;21(5):429-31.

223. Chacko J, Pramod K, Sinha S, Saini J, Mahadevan A, Bharath RD, Bindu PS, Yasha TC, Taly AB. Clinical, neuroimaging and pathological features of 5-nitroimidazole-induced encephalo-neuropathy in two patients: Insights into possible pathogenesis. Neurol India 2011;59:743-7.

224. Karbalay-Doust S, Noorafshan A. Ameliorative effects of curcumin on the spermatozoon tail length, count, motility and testosterone serum level in metronidazole-treated mice. Prague Med Rep. 2011;112(4):288-97.

225. Grover JK, Vats V, Srinivas M, Das SN, Jha P et al. Effect of metronidazole on spermatogenesis and FSH, LH and testosterone levels of pre-pubertal rats. Indian J Exp Biol. 2001 Nov;39(11):1160-2.

226. McClain RM, Downing JC, Edgcomb JE. Effect of metronidazole on fertility and testicular function in male rats. Fundam Appl Toxicol. 1989 Apr;12(3):386-96.

227. Pouly S, Storch MK, Matthieu JM, Lassmann H, Monnet-Tschudi F, Honegger P. Demyelination induced by protein kinase C-activating tumor promoters in aggregating brain cell cultures. J Neurosci Res. 1997 Jul 15;49(2):121-32.
228. Limpert AS, Bai S, Narayan M, Wu J, Yoon SO, Carter BD, Lu QR. NF-κB forms a complex with the chromatin remodeler BRG1 to regulate Schwann cell differentiation. J Neurosci. 2013 Feb 6;33(6):2388-97. doi: 10.1523/JNEUROSCI.3223-12.2013.

229. Medina PP, Sanchez-Cespedes M. Involvement of the chromatin-remodeling factor BRG1/SMARCA4 in human cancer. Epigenetics. 2008 Mar-Apr;3(2):64-8. Epub 2008 Apr 17.

230. Frensing T, Kaltschmidt C, Schmitt-John T. Characterization of a neuregulin-1 gene promoter: positive regulation of type I isoforms by NF-kappaB. Biochim Biophys Acta. 2008 Feb;1779(2):139-44. Epub 2007 Dec 3.

231. Johnson TA, Elbi C, Parekh BS, Hager GL, John S. Chromatin remodeling complexes interact dynamically with a glucocorticoid receptor-regulated promoter. Mol Biol Cell. 2008 Aug;19(8):3308-22. doi: 10.1091/mbc.E08-02-0123. Epub 2008 May 28.

232. Chan JY, Biden TJ, Laybutt DR. Cross-talk between the unfolded protein response and nuclear factor-κB signalling pathways regulates cytokine-mediated beta cell death in MIN6 cells and isolated mouse islets. Diabetologia. 2012 Nov;55(11):2999-3009. doi: 10.1007/s00125-012-2657-3. Epub 2012 Jul 28.

233. Volden PA, Conzen SD. The influence of glucocorticoid signaling on tumor progression. Brain Behav Immun. 2013 Mar;30 Suppl:S26-31. doi: 10.1016/j.bbi.2012.10.022. Epub 2012 Nov 16.

234. Widén C, Gustafsson JA, Wikström AC. Cytosolic glucocorticoid receptor interaction with nuclear factor-kappa B proteins in rat liver cells. Biochem J. 2003 Jul 1;373(Pt 1):211-20.

235. Kiyohara C and Yoshimasu K. Molecular epidemiology of major depressive disorder. Environ Health Prev Med. 2009 Mar;14(2):71-87. Epub 2009 Jan 20.

236. Kohli MA, Lucae S, Saemann PG, Schmidt MV, Demirkan A, Hek K et al. The neuronal transporter gene SLC6A15 confers risk to major depression. Neuron. 2011 Apr 28;70(2):252-65.

237. Koene S, Kozicz TL, Rodenburg RJ, Verhaak CM, de Vries MC, Wortmann S et al. Major depression in adolescent children consecutively diagnosed with mitochondrial disorder. J Affect Disord. 2009 Apr;114(1-3):327-32. Epub 2008 Aug 9.

238. Kolassa IT, Ertl V, Eckart C, Glöckner F, Kolassa S, Papassotiropoulos A et al. Association study of trauma load and SLC6A4 promoter polymorphism in posttraumatic stress disorder: evidence from survivors of the Rwandan genocide. J Clin Psychiatry. 2010 May;71(5):543-7. Epub 2010 Apr 6.

239. Müller MB and Wurst W. Getting closer to affective disorders: the role of CRH receptor systems. Trends Mol Med. 2004 Aug;10(8):409-15.

240. Jeremias J, Ledger WJ and Witkin SS. Interleukin 1 receptor antagonist gene polymorphism in women with vulvar vestibulitis. Am J Obstet Gynecol. 2000 Feb;182(2):283-5.

241. Witkin SS, Gerber S, Ledger WJ. Influence of interleukin-1 receptor antagonist gene polymorphism on disease. Clin Infect Dis. 2002 Jan 15;34(2):204-9. Epub 2001 Dec 7.

242. Witkin SS, Gerber S, Ledger WJ. Differential characterization of women with vulvar vestibulitis syndrome. Am J Obstet Gynecol. 2002 Sep;187(3):589-94.

243. Scarlett JM, Jobst EE, Enriori PJ, Bowe DD, Batra AK, et al. Regulation of central melanocortin signaling by interleukin-1 beta. Endocrinology. 2007 Sep;148(9):4217-25. Epub 2007 May 24.

244. Tkachenko IV, Jääskeläinen T, Jääskeläinen J, Palvimo JJ, Voutilainen R. Interleukins 1α and 1β as regulators of steroidogenesis in human NCI-H295R adrenocortical cells. Steroids. 2011 Sep-Oct;76(10-11):1103-15. Epub 2011 May 8.

245. Sasayama D, Hori H, Iijima Y, Teraishi T, Hattori K et al. Modulation of cortisol responses to the DEX/CRH test by polymorphisms of the interleukin-1beta gene in healthy adults. Behav Brain Funct. 2011 Jul 5;7:23. doi: 10.1186/1744-9081-7-23.

246. Wang X, Wu H, Miller AH. Interleukin 1alpha (IL-1alpha) induced activation of p38 mitogen-activated protein kinase inhibits glucocorticoid receptor function. Mol Psychiatry. 2004 Jan;9(1):65-75

247. Taves MD, Gomez-Sanchez CE, Soma KK. Extra-adrenal glucocorticoids and mineralocorticoids: evidence for local synthesis, regulation, and function. Am J Physiol Endocrinol Metab. 2011 Jul;301(1):E11-24. Epub 2011 May 3.

248. Vukelic S, Stojadinovic O, Pastar I, Rabach M, Krzyzanowska A, et al. Cortisol synthesis in epidermis is induced by IL-1 and tissue injury. J Biol Chem. 2011 Mar 25;286(12):10265-75. Epub 2011 Jan 14.

249. Slominski A, Zbytek B, Szczesniewski A, Semak I, Kaminski J, et al. CRH stimulation of corticosteroids production in melanocytes is mediated by ACTH. Am J Physiol Endocrinol Metab. 2005 Apr;288(4):E701-6. Epub 2004 Nov 30.

250. Adameyko I, Lallemend F. Glial versus melanocyte cell fate choice: Schwann cell precursors as a cellular origin of melanocytes. Cell. Mol. Life Sci. (2010) 67:3037–3055.

251. Dyer JK, Philipsen HL, Tonnaer JA, Hermkens PH, Haynes LW. Melanocortin analogue Org2766 binds to rat Schwann cells, upregulates NGF low-affinity receptor p75, and releases neurotrophic activity. Peptides. 1995;16(3):515-22.

252. Yamaura K, Doi R, Suwa E, Ueno K. A novel animal model of pruritus induced by successive application of glucocorticoid to mouse skin. J Toxicol Sci. 2011 Aug;36(4):395-401.

253. O'Keeffe GW, Gutierrez H, Pandolfi PP, Riccardi C, Davies AM. NGF-promoted axon growth and target innervation requires GITRL-GITR signaling. Nat Neurosci. 2008 Feb;11(2):135-42. Epub 2008 Jan 6.

254. Babula O, Linhares IM, Bongiovanni AM, Ledger WJ, Witkin SS. Association between primary vulvar vestibulitis syndrome, defective induction of tumor necrosis factor-alpha, and carriage of the mannose-binding lectin codon 54 gene polymorphism. Am J Obstet Gynecol. 2008 Jan;198(1):101.e1-4.

255. Grönwall C, Chen Y, Vas J, Khanna S, Thiel S, et al. MAPK phosphatase-1 is required for regulatory natural autoantibody-mediated inhibition of TLR responses. Proc Natl Acad Sci USA 2012 Nov 8. [Epub ahead of print]

256. Ábrahám IM, Harkany T and Luiten PGM. Action of Glucocorticoids on Survival of Nerve Cells: Promoting Neurodegeneration or Neuroprotection? Journal of Neuroendocrinology, 2001, Vol. 13, 749±760.

257. Yeager MP, Pioli PA,Guyre MP. Cortisol exerts bi-phasic regulation of inflammation in humans. Dose Response 2011;9(3):332-47. Epub 2010 Aug 12.

258. Erdeljan P, MacDonald JF, Matthews SG. Glucocorticoids and serotonin alter glucocorticoid receptor (GR) but not mineralocorticoid receptor (MR) mRNA levels in fetal mouse hippocampal neurons, in vitro. Brain Res. 2001 Mar 30;896(1-2):130-6.

259. Kino T, Manoli I, Kelkar S, Wang Y, Su YA, Chrousos GP. Glucocorticoid receptor (GR) beta has intrinsic, GRalpha-independent transcriptional activity. Biochem Biophys Res Commun. 2009 Apr 17;381(4):671-5. doi: 10.1016/j.bbrc.2009.02.110. Epub 2009 Feb 25.

260. Lewis-Tuffin LJ, Cidlowski JA. The physiology of human glucocorticoid receptor beta (hGRbeta) and glucocorticoid resistance. Ann N Y Acad Sci. 2006 Jun;1069:1-9.

261. Colomar A, Marty V, Combe C, Médina C, Parnet P, Amédée T.  [The immune status of Schwann cells: what is the role of the P2X7 receptor?]. [Article in French] J Soc Biol. 2003;197(2):113-22.

262. Meyer zu Hörste G, Hu W, Hartung HP, Lehmann HC, Kieseier BC. The immunocompetence of Schwann cells. Muscle Nerve. 2008 Jan;37(1):3-13.

263. Goethals S, Ydens E, Timmerman V, Janssens S. Toll-like receptor expression in the peripheral nerve. Glia. 2010 Nov 1;58(14):1701-9.

264. Ydens E, Lornet G, Smits V, Goethals S, Timmerman V, Janssens S. The neuroinflammatory role of Schwann cells in disease. Neurobiol Dis. 2013 Mar 21. pii: S0969-9961(13)00093-4. doi: 10.1016/j.nbd.2013.03.005. [Epub ahead of print]

265. Martini R, Fischer S, López-Vales R, David S. Interactions between Schwann cells and macrophages in injury and inherited demyelinating disease. Glia. 2008 Nov 1;56(14):1566-77. doi: 10.1002/glia.20766.

266. Chatterjea D, Wetzel A, Mack M, Engblom C, Allen J, et al. Mast cell degranulation mediates compound 48/80-induced hyperalgesia in mice. Biochem Biophys Res Commun. 2012 Aug 24;425(2):237-43. doi: 10.1016/j.bbrc.2012.07.074. Epub 2012 Jul 22.

267. Shumilina E, Zemtsova IM, Heise N, Schmid E, Eichenmüller M, Tyan L, Rexhepaj R, Lang F. Phosphoinositide-dependent kinase PDK1 in the regulation of Ca2+ entry into mast cells. Cell Physiol Biochem. 2010;26(4-5):699-706. Epub 2010 Oct 29.

268. Gaietta GM, Yoder EJ, Deerinck T, Kinder K, Hanono A, et al. 5-HT2a receptors in rat sciatic nerves and Schwann cell cultures. J Neurocytol. 2003 May;32(4):373-80.

269. Yoder EJ, Lee B, Ellisman MH. The expression of serotonin receptors by cultured rat Schwann cells is a function of their differentiation: correlation with a quiescent myelinating phenotype. Mol Cell Neurosci. 1997;8(5):303-10.

270. Dvorak AM, Morgan ES. Diamine oxidase-gold enzyme-affinity ultrastructural demonstration that human gut mucosal mast cells secrete histamine by piecemeal degranulation in vivo. J Allergy Clin Immunol. 1997 Jun;99(6 Pt 1):812-20.

271. Medic N, Lorenzon P, Vita F, Trevisan E, Marchioli A, et al. Mast cell adhesion induces cytoskeletal modifications and programmed cell death in oligodendrocytes. J Neuroimmunol. 2010 Jan 25;218(1-2):57-66. Epub 2009 Nov 10.

272. Tominaga M. [Activation and regulation of nociceptive transient receptor potential (TRP) channels, TRPV1 and TRPA1]. [Article in Japanese] Yakugaku Zasshi. 2010 Mar;130(3):289-94.

273. Fujita F, Uchida K, Moriyama T, Shima A, Shibasaki K et al. Intracellular alkalization causes pain sensation through activation of TRPA1 in mice. J Clin Invest. 2008 Dec;118(12):4049-57. doi: 10.1172/JCI35957. Epub 2008 Nov 13.

274. Johnson D, Seeldrayers PA, Weiner HL. The role of mast cells in demyelination. 1. Myelin proteins are degraded by mast cell proteases and myelin basic protein and P2 can stimulate mast cell degranulation. Brain Res. 1988 Mar 15;444(1):195-8.

275. Clarner T, Diederichs F, Berger K et al. Myelin debris regulates inflammatory responses in an experimental demyelination animal model and multiple sclerosis lesions. Glia. 2012 Jun 11. doi: 10.1002/glia.22367. [Epub ahead of print]

276. Yang XF, Wang H, Wen L. From myelin debris to inflammatory responses: a vicious circle in diffuse axonal injury. Med Hypotheses. 2011 Jul;77(1):60-2. Epub 2011 Apr 2.

277. Schweinhardt P, Kuchinad A, Pukall CF, Bushnell MC. Increased gray matter density in young women with chronic vulvar pain. Pain. 2008 Dec;140(3):411-9. Epub 2008 Oct 17.

278. Popoli M, Yan Z, McEwen BS, Sanacora G. The stressed synapse: the impact of stress and glucocorticoids on glutamate transmission. Nat Rev Neurosci. 2011 Nov 30;13(1):22-37. doi: 10.1038/nrn3138.

279. Sanacora G, Treccani G, Popoli M. Towards a glutamate hypothesis of depression: an emerging frontier of neuropsychopharmacology for mood disorders. Neuropharmacology. 2012 Jan;62(1):63-77. doi: 10.1016/j.neuropharm.2011.07.036. Epub 2011 Aug 3.

280. Musazzi L, Racagni G, Popoli M. Stress, glucocorticoids and glutamate release: effects of antidepressant drugs. Neurochem Int. 2011 Aug;59(2):138-49. doi: 10.1016/j.neuint.2011.05.002. Epub 2011 Jun 13.

281. Xiao L, Feng C, Chen Y. Glucocorticoid rapidly enhances NMDA-evoked neurotoxicity by attenuating the NR2A-containing NMDA receptor-mediated ERK1/2 activation. Mol Endocrinol. 2010 Mar;24(3):497-510. Epub 2010 Feb 16

282. Ruiz A, Matute C, Alberdi E. Intracellular Ca2+ release through ryanodine receptors contributes to AMPA receptor-mediated mitochondrial dysfunction and ER stress in oligodendrocytes. Cell Death Dis. 2010 Jul 15;1:e54. doi: 10.1038/cddis.2010.31.

283. Fu Y, Wang H, Huff TB, Shi R, Cheng JX. Coherent anti-Stokes Raman scattering imaging of myelin degradation reveals a calcium-dependent pathway in lyso-PtdCho-induced demyelination. J Neurosci Res. 2007 Oct;85(13):2870-81.

284. Fu Y, Sun W, Shi Y, Shi R, Cheng JX. Glutamate excitotoxicity inflicts paranodal myelin splitting and retraction. PLoS One. 2009 Aug 20;4(8):e6705. doi: 10.1371/journal.pone.0006705.

285. Alexander JK, DeVries AC, Kigerl KA, Dahlman JM, Popovich PG. Stress exacerbates neuropathic pain via glucocorticoid and NMDA receptor activation. Brain Behav Immun. 2009 Aug;23(6):851-60. Epub 2009 Apr 8.

286. Tse YC, Bagot RC, Wong TP- Dynamic regulation of NMDAR function in the adult brain by the stress hormone corticosterone. Front Cell Neurosci 2012;6:9. Epub 2012 Mar 6.

287. Kolber BJ, Montana MC,  Carrasquillo Y,  Xu J, Heinemann SF, Muglia LJ, Gereau RW. Activation of metabotropic glutamate receptor 5 in the amygdale modulates pain-like behaviour. J Neurosci. 2010 June 16; 30(24): 8203–8213.

288. Joéls M, Karst H. Corticosteroid effects on calcium signaling in limbic neurons. Cell Calcium. 2012 Mar-Apr;51(3-4):277-83. Epub 2011 Dec 6.
289. Werkman TR, Van der Linden S, Joëls M. Corticosteroid effects on sodium and calcium currents in acutely dissociated rat CA1 hippocampal neurons. Neuroscience. 1997 Jun;78(3):663-72.

290. Groeneweg FL, Karst H, de Kloet ER, Joëls M. Mineralocorticoid and glucocorticoid receptors at the neuronal membrane, regulators of nongenomic corticosteroid signalling. Mol Cell Endocrinol. 2012 Mar 24;350(2):299-309. Epub 2011 Jun 28.

291. Korte SM. Corticosteroids in relation to fear, anxiety and psychopathology. Neurosci Biobehav Rev. 2001 Mar;25(2):117-42.

292. Di  S,  Maxson MM, Franco A and Tasker JG. Glucocorticoids Regulate Glutamate and GABA Synapse-Specific Retrograde Transmission via Divergent Non-Genomic Signaling Pathways J Neurosci. 2009 January 14; 29(2): 393–401.

293. Bangasser DA. Sex differences in stress-related receptors: ″micro″ differences with ″macro″ implications for mood and anxiety disorders. Biol Sex Differ. 2013 Jan 21;4(1):2. doi: 10.1186/2042-6410-4-2.

294. Pariante CM. The glucocorticoid receptor: part of the solution or part of the problem? J Psychopharmacol. 2006 Jul;20(4 Suppl):79-84.

295. Nikisch G. Involvement and role of antidepressant drugs of the hypothalamic-pituitary-adrenal axis and glucocorticoid receptor function. Neuro Endocrinol Lett. 2009 Mar;30(1):11-6.

296. Iyo AH, Feyissa AM, Chandran A, Austin MC, Regunathan S, Karolewicz B. Chronic corticosterone administration down-regulates metabotropic glutamate receptor 5 protein expression in the rat hippocampus. Neuroscience. 2010 Sep 15;169(4):1567-74. Epub 2010 Jun 23.

297. Zunszain PA, Anacker C, Cattaneo A, Carvalho LA, Pariante CM. Glucocorticoids, cytokines and brain abnormalities in depression. Prog Neuropsychopharmacol Biol Psychiatry. 2011 Apr 29;35(3):722-9. Epub 2010 Apr 18.

298. Marin TJ, Martin TM, Blackwell E, Stetler C, Miller GE. Differentiating the impact of episodic and chronic stressors on hypothalamic-pituitary-adrenocortical axis regulation in young women. Health Psychol. 2007 Jul;26(4):447-55.

299. Jeckel CM, Lopes RP, Berleze MC et al. Neuroendocrine and immunological correlates of chronic stress in 'strictly healthy' populations. . Neuroimmunomodulation. 2010;17(1):9-18. Epub 2009 Oct 5.